ArticlePDF Available

Ongoing introgression of a secondary sexual plumage trait in a stable avian hybrid zone

Authors:

Abstract

Hybrid zones are dynamic systems where natural selection, sexual selection, and other evolutionary forces can act on reshuffled combinations of distinct genomes. The movement of hybrid zones, individual traits, or both are of particular interest for understanding the interplay between selective processes. In a hybrid zone involving two lek-breeding birds, secondary sexual plumage traits of Manacus vitellinus, including bright yellow collar and olive belly color, have introgressed asymmetrically ~50 km across the genomic center of the zone into populations more genetically similar to Manacus candei. Males with yellow collars are preferred by females and are more aggressive than parental M. candei, suggesting that sexual selection was responsible for the introgression of male traits. We assessed the spatial and temporal dynamics of this hybrid zone using historical (1989 - 1994) and contemporary (2017 - 2020) transect samples to survey both morphological and genetic variation. Genome-wide SNP data and several male phenotypic traits show that the genomic center of the zone has remained spatially stable, whereas the olive belly color of male M. vitellinus has continued to introgress over this time period. Our data suggest that sexual selection can continue to shape phenotypes dynamically, independent of a stable genomic transition between species.
Evolution, 2024, XX(XX), 1–15
https://doi.org/10.1093/evolut/qpae076
Advance access publication 16 May 2024
Original Article
Ongoing introgression of a secondary sexual plumage trait
in a stable avian hybrid zone
Introgresión continua de plumaje sexual secundario en una zona híbrida aviar estable
Kira M. Long1,2,, Angel G. Rivera-Colón3,4,, Kevin F. P. Bennett5,6,, Julian M. Catchen3,,
Michael J. Braun5,6,, Jeffrey D. Brawn7,
1Program in Ecology, Evolution and Conservation Biology, University of Illinois at Urbana-Champaign, Urbana, IL, United States
2Department of Fish and Wildlife Sciences, University of Idaho, Moscow, ID, United States
3Department of Evolution, Ecology, and Behavior, University of Illinois at Urbana-Champaign, Urbana, IL, United States
4Institute of Ecology and Evolution, University of Oregon, Eugene, OR, United States
5Behavior, Ecology, Evolution, and Systematics Program, University of Maryland, College Park, MD, United States
6Department of Vertebrate Zoology, National Museum of Natural History, Smithsonian Institution, Washington, DC, United States
7Department of Natural Resources & Environmental Sciences, University of Illinois at Urbana-Champaign, Urbana, IL, United States
Corresponding authors: 875 Perimeter Drive MS 1138, Moscow, ID 83844, United States. Email: kiralong778@gmail.com; W-523 Turner Hall, 1102 S. Goodwin
Avenue, Urbana, IL 61801, United States. Email: jbrawn@illinois.edu
Abstract
Hybrid zones are dynamic systems where natural selection, sexual selection, and other evolutionary forces can act on reshuffled combinations
of distinct genomes. The movement of hybrid zones, individual traits, or both are of particular interest for understanding the interplay between
selective processes. In a hybrid zone involving two lek-breeding birds, secondary sexual plumage traits of Manacus vitellinus, including bright
yellow collar and olive belly color, have introgressed ~50 km asymmetrically across the genomic center of the zone into populations more geneti-
cally similar to Manacus candei. Males with yellow collars are preferred by females and are more aggressive than parental M. candei, suggesting
that sexual selection was responsible for the introgression of male traits. We assessed the spatial and temporal dynamics of this hybrid zone
using historical (1989–1994) and contemporary (2017–2020) transect samples to survey both morphological and genetic variation. Genome-wide
single nucleotide polymorphism data and several male phenotypic traits show that the genomic center of the zone has remained spatially stable,
whereas the olive belly color of male M. vitellinus has continued to introgress over this time period. Our data suggest that sexual selection can
continue to shape phenotypes dynamically, independent of a stable genomic transition between species.
Keywords: manakin, RADseq, clines, hybridization, population genomics, sexual selection
Introduction
Hybridization is pervasive in nature and a key contributor to
speciation and diversication processes (Abbott et al., 2013).
Hybrid zones provide opportunities to explore how species
originate from diverging populations that may still be experi-
encing some degree of gene ow. Several speciation and diver-
sication processes can be observed in hybrid zones, including
reinforcement of barriers to admixture (Coyne & Orr, 2004;
Roberts & Mendelson, 2020), breakdown of reproductive
barriers allowing fusion between hybridizing species (Kearns
et al., 2018; Seehausen et al., 1997), creation of entirely new
forms through hybrid speciation (Gross & Rieseberg, 2005),
and adaptive introgression of traits from one species into
another. Other hybrid zones appear to be at least quasi- stable,
remaining in stasis for indenite periods of time (Pintoet al.,
2019; Wang et al., 2019); yet a closer study of such zones
may reveal dynamic equilibria between opposing evolution-
ary forces (Barton, 1979; Barton & Hewitt, 1985).
The movement or stability of hybrid zones is inuenced by
many different evolutionary factors. Moving hybrid zones
have been observed in several taxa and can move slowly or
many kilometers in a single year, especially if one parental
species has a selective or competitive advantage over the other
species (Dasmahapatra et al., 2002; Harr & Price, 2014;
Metzler et al., 2021; Reudink et al., 2007; Taylor et al., 2014;
Wang et al., 2014; Wielstra et al., 2017; Zohren et al., 2016).
Conversely, prolonged spatial stability is often seen in “ten-
sion zones” (Barton & Hewitt, 1985), where hybrids have
lower tness than parentals and the zones are localized in
specic regions coinciding with population density troughs or
ecological transition zones (Barton & Hewitt, 1985; Rosser
et al., 2014; Ruegg, 2008; Smith et al., 2013). Thus, hybrid
zones present dynamic systems governed by diverse factors,
and their mobility or stasis may change over time along with
changing evolutionary forces (Engebretsen et al., 2016; Wang
et al., 2019). Moreover, hybrid zones often involve conict-
ing evolutionary forces, such as sexual selection and natu-
ral selection, where selection for the traits of one form can
be opposed by generalized selection against hybrids. The
outcome of these pressures over time and space is a topic of
Received April 3, 2023; accepted May 14, 2024
Associate Editor: Scott Taylor; Handling Editor: Tim Connallon
Published by Oxford University Press for The Society for the Study of Evolution (SSE) 2024. This work is written by (a) US Government employee(s) and is
in the public domain in the US.
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
2Long et al.
active investigation (Buggs, 2007; Wielstra, 2019; Yang et al.,
2022). Characterizing phenotypic and/or genetic changes in
hybrid zones over time can elucidate how species boundaries
may go through periods of ux and allow us to capture the
dynamics of evolution within these populations.
Genome-scale analyses now provide the power to parse
evolutionary change in unprecedented detail, revealing cryp-
tic gene ow, demographic and selective history, and the
genetic underpinnings of key traits in speciation and adap-
tation (Abbott et al., 2016; Moran et al., 2021; Payseur &
Rieseberg, 2016; Taylor & Larson, 2019). For example,
evidence for the prevalence of ongoing gene ow during
speciation has expanded dramatically with advances in
genome sequencing (Chan et al., 2020; Trigo et al., 2013).
Additionally, modern genomics allows for the detailed char-
acterization of hybrid zone movement in real time and the
identication of loci responsible for introgressing traits
(Billerman et al., 2019; Ottenburghs et al., 2017; Semenov et
al., 2021; Wielstra, 2019).
Hybrid systems with historical genetic sampling are par-
ticularly informative for understanding spatial dynamics
in hybrid zones. One such system is a hybrid zone between
golden-collared manakins (Manacus vitellinus) and white-
collared manakins (Manacus candei) in western Panama.
Both are lek-breeding species, in which males compete for
mating opportunities by performing elaborate displays (Day
et al., 2021; Kirwan & Green, 2011), leading to strong sexual
selection on male traits (Snow, 2004). This hybrid zone was
rst characterized nearly 30 years ago by studying clinal vari-
ation in morphological and genetic traits that differ between
the parental forms along ~100 km region of a 570 km transect
(Brumeld et al., 2001; Parsons et al., 1993). Male collar and
belly color transition across this hybrid zone from a golden
yellow throat and olive belly in the golden-collared manakin
to a white throat and yellow belly in the white- collared
manakin (Figure 1) (Brumeld et al., 2001; Parsons et al.,
1993). The phenotypic transition for collar color occurred at
the Río Changuinola, the largest river in the region, where
birds on one side of the river had yellow collars and birds on
the other side had white collars (Parsons et al., 1993). This
phenotypic transition was displaced by about 50 km from the
cline centers for all genetic markers and several morphometric
traits (Brumeld et al., 2001). A similarly displaced transition
was observed for male belly color, although that cline was
not as sharp as the collar color cline (Brumeld et al., 2001;
Parsons et al., 1993). These displaced clines result in a geo-
graphic region where males resemble golden- collared manak-
ins in two secondary sexual plumage traits but are otherwise
similar to white-collared manakins genetically and morpho-
metrically (Brumeld et al., 2001; Parchman et al., 2013;
Parsons et al., 1993). Assays of male aggression (McDonald
et al., 2001) and female choice (Stein & Uy, 2006b) indicated
that yellow collars are advantageous over white collars, sug-
gesting that sexual selection was responsible for collar color
Panama
2
34
5
6
8
9
9.5 10
Golden-collared manakin (M. vitellinus) range
Hybrid (M. candei x vitellinus) range
Sampling site
White-collared manakin (M. candei) range
N
Río
Changuinola
0510 20
Kilometers
15
Figure 1. Map of the Manacus hybrid zone in Bocas del Toro, Panama. Dots represent sampling sites, numbered according to the system used
in Brumfield et al. (2001). The shaded regions denote the approximate ranges of each parental species and hybrids in Panama based on historical
sampling. The cartoon birds illustrate the major phenotypic differences between the parental species and hybrid populations.
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
Evolution (2024), Vol. XX 3
introgression. Belly color is correlated with collar color in
these birds (Brumeld et al., 2001) and thus may be directly
or indirectly under sexual selection as well.
Our overarching aims are to understand the spatial and
temporal dynamics of hybrid zones and how they may be
inuenced by a mating system that promotes strong sexual
selection. We ask whether a strong preference for traits from
one species over another can lead to the movement of an
entire hybrid zone. Under this scenario, can particular traits
become uncoupled from the rest of the genome? Do male sec-
ondary sexual plumage traits continue to introgress? We test
these questions in a hybrid zone where strong selection for
traits of one species has asymmetrically introgressed across
the hybrid zone, but no information exists on the stability of
the rest of the genome.
Here, we report the results of a longitudinal reassessment
of morphological and genetic variation across the Manacus
hybrid zone using a newly collected set of contemporary
(2017–2020) samples that closely replicate the historical
(1989–1994) transect. We characterized the genetic struc-
ture and admixture of the hybrid zone in both the histori-
cal and contemporary transects using a genome-wide set
of single nucleotide polymorphism (SNP) markers derived
from Restriction site-Associated DNA sequencing (RADseq).
We also estimated cline structure at both time points for
previously identied male phenotypic traits of interest. We
leveraged this multiyear sampling to explore the temporal
dynamics of this hybrid zone, directly comparing the two
datasets to provide empirical quantication for movement
and stability in this hybrid system.
Methods
Study species and sampling sites
Manacus vitellinus and M. candei (Pipridae) are frugivorous
birds that inhabit the understory of secondary forest or for-
est edge. M. vitellinus occurs from northern Colombia to
Panama, while M. candei ranges from southern Mexico to
western Panama. For our historical dataset, we used many of
the same specimens from the core transect collections stud-
ied by Parsons et al. (1993), Brumeld et al. (2001), Yuri et
al. (2009), and Parchman et al. (2013). We then revisited the
same sampling sites to collect contemporary samples. We rep-
licated the core transect sampling design of Brumeld et al.
(2001) as closely as possible, given land use changes in the
intervening years between historical sampling in 1989–1994
and contemporary sampling in 2017–2020. We sampled
Manacus populations at nine distinct sites in Bocas del Toro,
Panama, located an average of 11.56 km apart (Figure 1). We
used site numbers corresponding to the numbering scheme
used in Brumeld et al. (2001). Sampling site 9.5 (Miramar),
located between sites 9 and 10, was added to the contempo-
rary dataset to increase the spatial resolution of geographic
cline analysis (see Methods below) and did not have previ-
ous historical transect data. We chose to focus our sampling
within the active hybrid zone area, so sampling sites outside
the province of Bocas del Toro were excluded, along with site
7 which deviated the farthest from the transect line. The nine
sampling sites represent two populations of M. candei, two
populations of M. vitellinus, and ve populations of hybrids
in the known transition zone (Figure 1; Table 1).
Field and museum data collection
We used 36 mm mesh “ATX” type mist nets on leks to capture
individuals. Each bird was given a numbered aluminum band
and a combination of plastic-colored leg bands to facilitate
individual identication. Birds were examined to determine
age and sex using eld observations of denitive adult male
plumage and the presence of a brood patch for reproduc-
tive females. Nonbreeding females and juvenile males were
sexed using a molecular sexing protocol (see below). We col-
lected blood samples (<100 µl) in glass capillary tubes from
all individuals by puncturing the brachial vein. Blood was
stored in Longmire’s lysis buffer (Longmire et al., 1997) and
kept frozen to preserve long-term DNA quality. From wild-
caught males with denitive adult plumage, we measured
beard length (measured as the longest feather directly below
the front of the eye), epaulet width (measured as the widest
point of the white or yellow coverts from the bend of the
folded wing to the posterior edge of the white or yellow epau-
let feathers), and tail length (measured as the longest feather
in the tail). These three morphological traits were identied
as clinal in Brumeld et al. (2001). We remeasured these same
three male phenotypic traits for the historical dataset on
museum specimens at the Smithsonian National Museum of
Natural History (NMNH). In order to minimize observer bias
between the historical and contemporary datasets, all mea-
surements reported here were taken by KML.
We characterized male plumage colors using Ocean Insight
reectance spectrophotometers (USB2000+ in 2017 and
FLAME-S-UV-VIS-ES in 2018–2022). Both spectrophotom-
eters used a PX-2 pulsed xenon lamp (Ocean Insight). We
used a ber optic probe tted with a plastic probe holder
and took the feather color reads at a 90° angle. The probe
Table 1. Sampling sites and number of sequenced individuals retained for genetic analyses after quality filtering.
Sampling site number Location name Distance from site 10 (km) Putative hybrid status Historical N Contemporary N
2 San San Drury 92.50 M. candei 31 49
3 El Silencio 79.00 M. candei 19 26
4 Finca Cuatro 71.25 Hybrid 18 24
5 Río Oeste 48.50 Hybrid 17 20
6 Quebrada Pastores 42.50 Hybrid 4 24
8 Río Uyama 29.50 Hybrid 22 27
9 Palma Real 20.75 Hybrid 23 65
9.5 Miramar 8.44 M. vitellinus NA 30
10 Rambala 0.00 M. vitellinus 18 53
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
4Long et al.
holder minimized the ambient light in reectance measure-
ments and maintained the same distance between the probe
and the feathers for each read. The percent reectance across
all wavelengths was measured relative to a pure white reect-
ing standard (WS-1-SL Diffuse Reectance Std, Spectralon)
as 100% reectance and a 0% dark standard. The white
standard was tted with a plastic holder that kept the probe
from directly touching the white substrate of the standard,
keeping the white standard clean and free of discolorations.
Due to the added distance between the probe holder and the
white standard substrate, we calibrated our measurements
to account for the drop in reectance from this gap, which
would be present evenly across the entire spectrum on each
spectrophotometer. We adjusted the spectral curve values
using a 0.7 and 0.5 multiplier for the USB2000+ and FLAME-
S-UV-VIS-ES spectrophotometers, respectively, based on the
percent drop in white reectance as measured on both spec-
trophotometers following the calibration methods of Cook
et al. (2013), Kelly et al. (2012), and Troy Murphy (personal
communication). To assess bias in the data due to using two
different spectrophotometers, we plotted the reads taken by
each spectrophotometer and did not observe shifts in reads
based on which spectrophotometer was used (Supplementary
Figure S1; Supplementary Table S1).
The reectance data for the contemporary birds were taken
on live birds in the eld, while the reectance data for the
historical specimens were measured from museum skins at
the Smithsonian NMNH using the FLAME-S-UV-VIS-ES
spectrophotometer. Previous studies have found that the
plumage of museum specimens can change over time but that,
overall, the reectance spectra from well-preserved museum
specimens are similar to wild birds (Doucet & Hill, 2009).
Additionally, museum specimen degradation will depend on
how well-preserved the specimens are, the age of the speci-
men, and the mechanism underlying the color of interest. The
Manacus specimens at the Smithsonian NMNH used in this
study are between 31 and 26 years old and are clean and
in good condition. We took reectance spectra of the den-
itive male collar and belly plumage color near the center of
the throat or belly on each bird. Five spectra were taken for
each plumage patch, repositioning the probe between each
read and averaging across spectra for each bird. Following
Brumeld et al. (2001), we used the average adjusted reec-
tance at 490 nm and 665 nm to represent collar color and
belly color, respectively, because these wavelengths had the
greatest differences in mean reectance between the two
parental samples. Outlier readings beyond two standard devi-
ations away from the mean for any individual were removed
from the dataset. We used the average percent reectance at
each wavelength for geographic cline analysis.
All protocols involving live birds were reviewed and app-
roved by the Illinois Animal Care and Use Committee and the
Smithsonian Tropical Research Institute Animal Care and Use
Committee (Illinois IACUC numbers: 15234 & 18239; STRI
ACUC numbers: 2016-0301-2019 & 2019-0115-2022).
RADseq library preparation and genotyping
DNA from the contemporary 2017–2020 samples were
extracted from whole blood using the animal tissue proto-
col on a Gene Prep automated extractor (Autogen), which
involves Proteinase K digestion, phenol extraction, and alco-
hol precipitation. DNA concentrations were determined using
a Quant-iT dsDNA Broad-Range Assay Kit (Thermo Fisher
Scientic) on a SpectraMax iD3 plate reader running SoftMax
Pro 7 software. DNA stocks for the historical 1989–1994
samples were available from previous studies (Brumeld et al.,
2001; Parsons et al., 1993). All samples are deposited at the
NMNH, Smithsonian Institution, in Washington, DC. DNA
from individuals of unknown sex were assayed in an avian
polymerase chain reaction (PCR) molecular sexing protocol
(Fridolfsson & Ellegren, 1999). Briey, we made a master mix
using the primers 2550F = 5ʹ-GTT ACT GAT TCG TCT ACG
AGA-3ʹ and 2718R = 5ʹ-ATT GAA ATG ATC CAG TGC
TTG-3ʹ and TaKaRa ExTaq DNA Polymerase (Takara Bio
USA, Inc.). We ran the samples for 31 cycles and ran the PCR
product on a 2% agarose gel.
Six single-digest SbfI RADseq libraries were prepared using
the protocols described in Baird et al. (2008) and Etter et al.
(2011). After diluting all samples in the library to the same
concentration, 600–800 ng of genomic DNA per sample
(depending on the library, see Supplementary Tables S2 and
S3) was digested using the SbfI-HF enzyme (New England
Biolabs). Custom P1 adapters containing unique 7-bp bar-
codes (Hohenlohe et al., 2012) were ligated to each individual
sample with T4 ligase (New England Biolabs). The DNA from
all uniquely barcoded individuals in a library was pooled at
equimolar concentration, and 1500 ng of total DNA was then
sheared using Covaris (Covaris, Inc.) or Qsonica (Qsonica
LLC) sonicators and size selected for 300–600 bp inserts
using AMPure XP beads (Beckman Coulter). The sheared
DNA was end-repaired and ligated to custom P2 adapters.
The P2-ligated DNA was then amplied for 12 rounds of
PCR using custom RAD primers (Hohenlohe et al., 2012) and
cleaned with 0.8× AMPure XP beads (Beckman Coulter). The
nal libraries were sequenced either on an Illumina HiSeq
4000 or an Illumina NovaSeq 6000 to generate 2 × 150 bp
reads (Supplementary Table S3).
In total, 593 samples were processed, which included
542 unique individuals and 51 samples sequenced in dupli-
cate to ameliorate low coverage. A breakdown of the
detailed sequencing information per-library is provided in
Supplementary Table S3. Using the Stacks software version
2.60 (Rochette et al., 2019), raw reads were ltered with
process_radtags to remove reads with low quality (-q)
and uncalled bases (-c), rescue ambiguous barcodes (-r),
and demultiplex samples based on their unique barcode (-b).
For each library, we retained between 668 million to 1.6 bil-
lion reads (Supplementary Table S3).
The processed reads for all 593 sequenced samples were
mapped to the M. vitellinus reference assembly (NCBI RefSeq
Accession: GCF_001715985.3) using BWA mem version
0.1.17 (H. Li & Durbin, 2009; Heng Li, 2013) with default
parameters. Alignments were processed and sorted using
SAMtools version 1.7 view and sort (H. Li et al., 2009).
RAD loci assembly and genotyping from the processed align-
ments were done with the Stacks gstacks module version
2.60 (Rochette et al., 2019), where we removed PCR dupli-
cates (--rm-pcr-duplicates). In total, we assembled
426,521 raw RAD loci with an average nonredundant cov-
erage of 14.0× (SD 6.9×). To provide a baseline level ltering
of missing data, we ran the Stacks populations module
on all samples with an average per-locus coverage at 6× or
higher, requiring a minimum minor allele count of 3 (--mac
3), requiring a locus to be present in at least 80% of samples
per population (i.e., sampling sites) (-r 0.80), and be in all
9 populations (-p 9). Additionally, we used the --hwe ag
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
Evolution (2024), Vol. XX 5
to calculate differences between the observed and expected
genotype frequencies under Hardy–Weinberg equilibrium
(HWE) for each population. This ltered run kept 77,551 loci
and 584,680 variant sites across the 470 individuals that were
retained in the dataset (Table 1). This populations run is here-
after referred to as the “master run,” and its output was used
to create all whitelists of loci/SNPs for downstream analyses.
Population structure analysis
We used ADMIXTURE (Alexander et al., 2009) to assess pop-
ulation genetic structure in both the historical and contem-
porary datasets. Using a custom Python script, we created a
whitelist of 10,000 SNPs sampled from our master popula-
tions run. First, we ltered the data from the master run for
SNPs with a minimum allele frequency of 1% and in HWE in
all populations and selected a single, randomly chosen SNP at
each locus in order to ensure the independence of markers in
the dataset. From this ltered subset, the script then randomly
selects 10,000 sites to generate a nal whitelist. For each data-
set, we reran the Stacks populations module, supplying the
whitelist and exporting genotypes in the PLINK (--plink)
and GENEPOP (--genepop) format. The resulting .ped
les were converted to binary .bed les using PLINK version
1.90b6.25 (Chang et al., 2015) and ADMIXTURE version
1.3.0 was run separately on the historical and contemporary
datasets for population cluster (K) values 1–9. To nd an
optimal value for the number of population clusters (K), we
followed the ADMIXTURE documentation by plotting the
cross-validation error for each K value. To corroborate the
results of ADMIXTURE using an ordination method instead
of a clustering method, we ran a principal component anal-
ysis (PCA) using the same whitelisted SNP dataset described
above in the ADMIXTURE analysis. We ran the PCA in ade-
genet version 2.1.7, an R package for multivariate analysis of
genetic markers (Jombart, 2008; Jombart & Ahmed, 2011).
The ADMIXTURE and PCA results were plotted in R version
4.2.1 (R Core Team, 2022).
Hybrid index
In order to compare the shifts in the location of the genetic
center of the hybrid zone over time, we calculated the hybrid
index for each of 470 individuals across historical and con-
temporary datasets using the R package gghybrid version
1.0.0 (Bailey, 2020), which uses a Bayesian Markov chain
Monte Carlo (MCMC) method for estimating the hybrid
index. First, using the same whitelist of 10,000 SNPs as the
ADMIXTURE analysis, we reran populations to export
the genotypes in the STRUCTURE format (--structure).
Using the gghybrid data input function (data.prep()),
we assigned the parental source populations 0 and 1 as M.
candei (sampling site 2) and M. vitellinus (sampling site 10),
respectively. Using the max.S.MAF option, we removed loci
for which the smaller of the two parental minor allele fre-
quencies is greater than 10%, ensuring that all SNPs used in
the hybrid index are informative by requiring the allele fre-
quency of at least one parental population to be close to 0.
This lter also allows for the retention of SNPs that are not
xed for alternate alleles in parental populations, providing
for denser sampling of the entire genome in estimating the
hybrid index. We calculated the Bayesian hybrid index using
the esth() gghybrid function, running for a total of 50,000
iterations with 10,000 burn-ins. This analysis was performed
separately for the historical and contemporary datasets. We
recorded the mean hybrid index value for each individual and
used the mean for each sampling site in downstream analyses.
Assignment of hybrid class
We classied our 470 Manacus individuals among different
hybrid types (i.e., if an individual is an F1, F2, backcross,
etc.) by establishing the relationship between the parental
proportion (in the form of hybrid index) and their interspe-
cic heterozygosity. The interspecic heterozygosity statis-
tic calculates the proportion of an individual’s genome that
contains alleles inherited from both parental populations. To
establish this relationship, we rst obtained a subset of 2,000
diagnostic alleles between the two parental populations (see
Supplementary Methods). Diagnostic alleles were used to
estimate interspecic heterozygosity, which can be sensitive
to alleles of intermediate frequencies present in the parental
populations (Fitzpatrick, 2012).
Using the genotypes of these 2,000 diagnostic markers, we
rst reestimated the hybrid index of the 470 individuals using
the methods described above (using the esth() function in
gghybrid). Then, we calculated interspecic heterozygosity
using the R package Introgress version 1.2.3 (Gompert &
Buerkle, 2010). In Introgress, we rst processed the genotype
data for both parental and admixed individuals using the
prepare.data() function, dening the input genotypes as
co-dominant markers (coded as “C in the loci.data le)
and not xed in the parental populations (fixed=FALSE).
The interspecic heterozygosity of each individual was then
estimated using the calc.intersp.het() function. We
used “triangle plots” to show the interspecic heterozygosity
as a function of the hybrid index (Supplementary Figure S4),
classifying individuals among different hybrid types based on
their placement between the two axes.
The hybrid classication analysis was additionally repeated
on a simulated dataset. Multigeneration hybrid genotypes
were simulated from the pool of parental genotypes obtained
from the 2,000 diagnostic SNPs. Our simulations were per-
formed using a custom Python script (see Supplementary
Methods), using a similar approach to the methods described
by (Lavretsky et al., 2016; Lavretsky et al., 2019; Wringe et
al., 2017). Briey, we used the empirical parental genotypes
to calculate the observed allele frequency across our 2,000
diagnostic SNPs. Then, we dened a series of dened crosses,
each describing the generations of hybridization (or back-
cross), the parents of each cross, as well as the genomic pro-
portions (Fitzpatrick, 2012; Turelli & Orr, 2000) expected.
Lastly, the parental allele frequencies and genomic propor-
tions were then used to determine the probability of geno-
types in the individuals of a given hybrid cross. The genotypes
of simulated individuals of known hybrid assignment were
then used to calculate hybrid index and interspecic heterozy-
gosity (Supplementary Figure S4), which were then compared
against the values observed in our Manacus empirical data.
Geographic clines
For the geographic clines, we reltered the master dataset to
obtain all sites that were in HWE in each of the 9 populations
and only a single SNP per locus, retaining a total of 72,506
loci. Sites in HWE were used for the geographic cline analysis
because sites that drastically depart from equilibrium could
bias cline ttings (Derryberry et al., 2014; Macholán et al.,
2007, 2008; Phillips et al., 2004). All geographic cline analy-
ses were repeated with and without the newly added sampling
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
6Long et al.
site 9.5. With the new 72.5K SNP whitelist, we reran the
Stacks populations module to export genotypes in the
HZAR format (--hzar). HZAR is an R package designed
to t geographic clines to one-dimensional transect data in
hybrid zones using the Metropolis-Hastings MCMC algo-
rithm (Derryberry et al., 2014). We wrote a custom R script
to parallelize HZAR version 0.2-5, allowing us to efciently
run the analysis independently for each selected SNP. Of the
72.5K whitelisted markers, we removed sites that were invari-
ant in either the contemporary or historical datasets, ensuring
that only sites that were variant in both datasets were run in
HZAR. In total, we retained 68,684 and 63,444 SNPs in the
contemporary and historical datasets, respectively. This rela-
tively large number of markers was retained for each dataset
in order to maximize sampling of clinal variation across the
genome. For each SNP in the historical and contemporary
datasets, we ran HZAR using four cline models (as originally
described by Szymura and Barton (1986, 1991)) to maintain
consistency with the analyses performed by Brumeld et al.
(2001). The four models were: (1) a model with the observed
minimum (pmin) and maximum (pmax) allele frequencies and
without tting exponential decay curves for the cline tails, (2)
a model with estimated pmin and pmax without tted exponen-
tial decay curves, (3) a model with estimated pmin and pmax and
tting decay curves to both cline tails, and (4) a null model
without t for either allele frequency or cline tails. For each
model, we ran three independent chains composed of three
dependent model t requests, each of 100,000 length and with
10,000 burn-in generations. We randomly selected a subset
of the genetic clines to inspect the MCMC trace plots of the
chains to verify that the models were converging to a local
optimum. The best-t model was selected using the Akaike
Information Criterion with correction for small sample size
(AICc). We then extracted the estimated cline width and cline
center values with 95% Bayesian Credibility Intervals (CI) for
the best model of each cline, along with the maximum likeli-
hood cline generated by the best model. We removed sites that
were assigned null models (i.e., indicating that the SNP was
not clinal and thus contained null values for the cline center
and width). In the historical and contemporary datasets, there
were 11,925 and 8,254 sites, respectively, that were nonclinal
and thus were assigned null models. We used a custom Python
script to combine the cline parameter outputs into a single
le and to select clines at loci that were present in both the
historical and contemporary datasets. The distribution of
all 48,316 cline centers is shown in Supplementary Figure
S2. Next, we removed SNPs displaying cline centers located
beyond the geographical boundaries of the studied area (i.e.,
cline centers less than 0 km or greater than 92.5 km). In total,
we retained cline results for 46,734 SNPs that passed all the
above lters and were observed in both historical and con-
temporary datasets (comprising our base ltering scheme).
Additionally, we applied subsequent lters based on the dif-
ference in allele frequencies between the parental sampling
sites and the range of the estimated CI, to further validate
our detection of movement between temporal datasets (see
Supplementary Methods). We considered a cline as showing
evidence of movement if the cline center CI did not overlap
between historical and contemporary datasets.
We also ran HZAR on ve male phenotypic traits known
to differ between M. vitellinus and M. candei: beard length,
epaulet width, tail length, collar color, and belly color. We
ran three morphological cline models: (1) a model with the
observed morphological data (“xed”) and without tting
exponential decay curves for the cline tails, (2) a model with
estimated “free” without tted exponential decay curves, (3)
a model with estimated “free” and tting decay curves to
both cline tails. After running each model three times with
a chain length of 100,000 and burn-in of 10,000, the best-
t model was selected using AICc. The cline analyses for the
contemporary dataset were performed both with and with-
out data for site 9.5 to directly match the historical dataset.
For any phenotypic traits that displayed substantial changes
in cline center or width estimates, we further validated these
results by running longer chain lengths of 1,000,000 with
100,000 burn-in and visually inspecting the MCMC trace
plots. In addition to the ve male phenotype clines, we ran a
geographic cline on the genomic hybrid index calculated by
gghybrid to assess the genome-wide movement of the hybrid
zone. We calculated the mean hybrid index values in each
sampling site for both the contemporary and historical data-
sets and analyzed them using HZAR. As with the SNP data,
we obtained the best cline model and exported the associated
center, width, and CI values. For both the phenotypic clines
and hybrid index clines, we assessed whether the clines had
moved based on whether the historical and contemporary CIs
for the cline centers overlapped.
Results
Genome-wide patterns of genetic population
structure are stable through time
We sampled 10,000 SNPs from 470 sequenced individu-
als spanning the historical and contemporary datasets for
admixture analyses. Overall, we found genetic population
structure across the transect to be consistent over time (Figure
2A). When we assessed admixture with two genetic clusters
(K = 2), individuals separated based on apparent parental spe-
cies ancestry. The genomic transition from M. vitellinus to M.
candei was centered near site 9, with substantially admixed
individuals present only at sites 8, 9, and 9.5. Although all
birds in sites 4–8 closely resemble M. vitellinus phenotypi-
cally, their genomes are M. candei-like, aside from a few inter-
mediate individuals at site 8.
K = 3 was the best-t model for both historical and con-
temporary datasets according to cross-validation analyses
(Supplementary Table S4). With K = 3, a third genetic clus-
ter separated hybrids with M. candei-like genomes east of
the Río Changuinola (sites 4–9) from M. candei to the west
(sites 2 and 3), highlighting the river as an important bar-
rier to gene ow (Figure 2A). The largest variance in admix-
ture proportions is still seen in sampling site 9 in both time
points. Patterns of admixture are consistent at K = 2 and
K = 3 clustering for the historical and contemporary tran-
sect samples, showing that overall admixture and genetic
clustering in the hybrid zone are stable over the time period
examined.
The PCA corroborated the admixture analysis (Figure
2B). Individuals in the PCA group by sampling site and
not by time also imply temporal stability in genome-wide
structure. The first principal component axis (PC1) sep-
arates individuals based on apparent parental ancestry,
explaining 14.54% of SNP variance. We found a tight cor-
relation between PC1 scores and genomic hybrid index,
suggesting that the primary axis of SNP variation strongly
corresponds with parental ancestry (Figure 2D). Sampling
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
Evolution (2024), Vol. XX 7
site 9, again, showed the largest variation of all the sam-
pling sites, spanning between the M. vitellinus parental
cluster and the other hybrid sites. Additionally, paren-
tal M. candei sites 2 and 3, both to the west of the Río
Changuinola, clustered tightly together and are separated
from the rest of the hybrid sites along PC2; however, this
axis only explains a small proportion of the observed vari-
ation (1.56%).
The genomic hybrid index reflects stable patterns
of parental ancestry
The distributions of hybrid indices were stable over time
and showed no substantial movement of the principal
genomic transition in the hybrid zone (Figure 2C). The geo-
graphic clines for the genomic hybrid index had very simi-
lar cline centers and cline widths with largely overlapping
CI from the historical and contemporary transect datasets
(Supplementary Table S5). The cline center CI from both
datasets overlapped the location of hybrid sampling site 9,
about 20.75 km from M. vitellinus parental site 10. Thus, the
genomic center of the hybrid zone has remained stable over
time near sampling site 9.
Birds sampled at sites 4–8 for both contemporary and his-
torical transects possessed a minor amount of M. vitellinus
ancestry in a majority M. candei genomic background, with
the exception of a single individual in contemporary site 8
with a hybrid index of 0.78 and a single individual in histor-
ical site 8 with a hybrid index of 0.50 (Supplementary Figure
S3; Supplementary Table S6). The contemporary samples
from newly assayed site 9.5 showed a majority M. vitellinus
ancestry, with a mean hybrid index of 0.9470 (SD 0.0612).
While extensive admixture was observed at site 9 in both the
contemporary and historical datasets, we did not nd evi-
dence of early generation hybrids in either the contemporary
or historical samples (Supplementary Figure S4). Hybrids
from site 9 had the highest levels of variation in the hybrid
index and interspecic heterozygosity in both the contempo-
rary and historical datasets (Supplementary Figures S3 and
S4; Supplementary Table S6). This result is consistent with the
population level structure analyses above, showing that site 9
is consistently the most variable and admixed locality in the
hybrid zone. Thus, the proportion of genome-wide parental
ancestry in hybrid individuals appears stable over time at the
population level.
Río
Changuinola
Contem-
porary
Historical
K = 2K = 3
Historical
Sampling Site
Manacus
vitellinus
Manacus
candei
0310 9.5090806 05 04 02
Collar
Belly
Contem-
porary
PC1 14.54%
PC2 1.56%
Historical
Contemporar
y
Sampling Site
10
4
9.5
9
6
5
8
3
2
Hybrid Index
PC1 14.54%
−50 050
0.00
0.25
0.50
0.75
1.00
Historical
Contemporar
y
Sampling Site
10
4
9.5
9
6
5
8
3
2
Km from site 10 (M. vitellinus)
020 40 60 80 100
+++
+++
+
++
++
+
+
+
+
++++++++++++++++++++++
++++++++
++++
+++
+
+
+
+
++++
+
++++++
+++
++
++
++
++
+++++++++
++++++++++++++++++++++++
Historical
Contemporary
+
+
Hybrid Index
0310 9.5 09 08 06 05 04 02
1.00
Sampling Site
0.80
0.60
0.40
0.20
0.00
AB
D
C
Figure 2. Patterns of population structure across the Manacus hybrid zone. (A) Admixture analysis of contemporary and historical manakin datasets
based on 10,000 SNPs. Sampling site numbers are listed at the bottom of the figure, with the M. vitellinus site 10 on the left and M. candei site 2 on
the right. Site 9.5 was sampled for the contemporary but not the historical dataset. Each vertical bar represents one bird; the width of contemporary
and historical population blocks was equalized for ease of comparison. The vertical dashed line between sites 3 and 4 represents the Río Changuinola.
The yellow, green, and white top bars provide a general, qualitative description of the predominant collar and belly color of males at each sampling
site, highlighting the phenotypic transition of plumage color at the Río Changuinola. (B) Principal component analysis based on 10,000 SNPs from
contemporary and historical manakin datasets. (C) Geographic clines of the hybrid index and associated 95% credibility intervals, with the historical
dataset in blue and contemporary in red. A hybrid index of 0 indicates complete M. candei ancestry (site 2), while a value of 1 indicates complete
M. vitellinus ancestry (site 10). The top x-axis shows the location of the 9 sampling sites. The crosses denote the mean hybrid index at each site. (D)
Correlation between the first principal component and genomic hybrid index.
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
8Long et al.
Genome-wide geographic clines have remained
stable
Of the 72,506 SNPs in the full RADseq dataset, we retained
46,734 SNPs that displayed clinal variation across the tran-
sect and passed all baseline lters for the cline center anal-
ysis. The overall distribution of contemporary clines with
and without the newly added sampling site 9.5 are shown
in Supplementary Figure S2, but the inclusion of the addi-
tional site did not largely change the cline center distribution.
The clines from the historical samples had a median cline
center at 23.85 km from site 10, the M. vitellinus parental
site (mean = 32.98 km, SD = 21.80 km). The contemporary
dataset had a median cline center at 22.75 km from site 10
(mean = 30.81 km, SD = 19.41 km). The average cline center
for both datasets appears to be largely robust to the effects
of ltering, with the median cline centers estimated between
22.44 and 23.85 km for both datasets across all ltering
schemes (Supplementary Table S7).
Both the historical and contemporary median cline centers
are between sampling sites 8 and 9, located at 29.50 km and
20.75 km from site 10, respectively. Similarly, 51.25% and
50.42% of all contemporary and historical clines, respec-
tively, have estimated cline centers between these two sam-
pling sites. While the specic proportion of cline centers
located between these two sampling sites varies depending
on the lters applied (ranging from 50.42% to 97.95%;
Supplementary Table S7), all ltering schemas nd the
majority of cline centers for the historical and contemporary
datasets located in the 8.75 km between sampling sites 8 and
9. This distribution of cline centers surrounding sampling
site 9 provides support for the assignment of this site as the
genomic center of the Manacus hybrid zone, in accordance
with previous ndings.
While, on average, cline centers are located proximally
to the previously characterized hybrid center (sampling site
9), we do nd evidence of displaced clines where the shift in
allele frequency occurs near the Río Changuinola (75.13 km
away from sampling site 10). Under the base ltering scheme,
we detected 3,166 (6.77%) clines with maximum likelihood
cline center estimates at or past the river in the contemporary
dataset and 4,421 (9.46%) in the historical dataset. While
the specic proportion of clines with centers at or past the
river is dependent on the effects of ltering (ranging from 0
to 9.46%; Supplementary Table S7), we do detect the dis-
placement of a small but measurable proportion of clines
away from the genetic center of the hybrid zone. For both
the historical and contemporary datasets, these clines with
centers along the Río Changuinola are likely examples of loci
introgressing along the hybrid zone, shifting from the genetic
hybrid center towards the current phenotypic transition.
We detected only 795 (1.70%) out of our total 46,000
SNPs from our base ltering scheme that exhibited evidence
of movement over time, showing nonoverlapping cline cen-
ter CIs between the two temporal datasets. The proportion
of clines centers with nonoverlapping CIs ranged between
0.77% and 3.66% across all ltering schemas (Supplementary
Table S7). From the base ltering scheme, the 795 putatively
moving clines are distributed across 91 different scaffolds in
the M. vitellinus reference assembly (Supplementary Table
S8). At our most stringent ltering scheme, ltering by 80%
allele frequency differences between the two parental pop-
ulations to observe a subset of more putatively diagnostic
alleles (see Supplementary Methods), we detect only six clines
with evidence of movement located in six different scaffolds
(Supplementary Tables S7 and S8). Since the M. vitellinus ref-
erence assembly is not chromosome-level, we used conserved
Beard Length
Belly Color
Collar Color
Hybrid Index
Tail Length
20 40 60 80
Estimated Cline Centers
Distance from M. vitellinus (km)
ACB
ED F
020406080100
0
20
40
60
80
100
Belly Color
% Reflectance (665 nm)
++++
++
+
+
+
+
+
+
+
+
+
++
+
+
+
+
+
+
+
++
+
+
+
+
+
++
++
++
+
+
+
+
+
+
+
+
+
+
+
+
+
++
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
020406080 100
0
20
40
60
80
100
Collar Color
% Reflectance (490 nm)
++++
++
++
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
++
++
+
+
+
+
+
+
+
++
Tail Length
020406080 100
20
25
30
35
40
Length (mm)
+
+
++++++
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
020406080100
0
2
4
6
8
10
12
14
Epaulet Width
Width (mm)
+
+
++
++
++
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
++
+
+
+
+
+
++
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
020406080 100
8
10
12
14
16
18
20
Beard Length
Length (mm)
+
+
+
+++++
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
Figure 3. Contemporary and historical clines were estimated from manakin phenotypic data across 8 sampling sites. For all panels, blue denotes the
historical dataset, red the contemporary dataset, and all x-axes are in kilometers from the parental M. vitellinus (site 10). Crosses in (A)–(E) represent
the average value of the phenotypic trait of interest in each sampling site. Shaded regions represent 95% credibility intervals for each cline. Purple
patches on the manakin cartoons highlight the plumage patches measured for each trait. (F) The location along the transect of the maximum likelihood
estimated cline center (dot) with 95% credibility intervals (horizontal bars) for the traits of interest. Hybrid index values pertain to the genomic hybrid
index cline shown in Figure 2C. The vertical dashed line represents the location of the Río Changuinola.
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
Evolution (2024), Vol. XX 9
synteny to identify orthologous chromosomes with the lance-
tailed manakin (Chiroxiphia lancaeolata) reference assembly
(see Supplementary Methods) to verify that the SNPs with evi-
dence of movement are on several separate chromosomes and
not clustered in a single region of the genome (Supplementary
Table S8). These results indicate that most genomic loci have
remained geographically stable over 30 years but that limited
cline movement has occurred.
Plumage introgression has advanced in 30 years
In addition to the distribution of allele frequencies across
space, we also t sigmoid clines to the observed variation
in ve male plumage traits (beard length, epaulet width, tail
length, collar color, and belly color) in both historical and
contemporary transect samples (Supplementary Tables S9 and
S10). Clines for four plumage traits are similar in the histori-
cal and contemporary transects (Figure 3A–D; Supplementary
Table S5), suggesting these traits are largely geographically
stable through time. Beard length and tail length cline cen-
ters lie near the genomic center of the hybrid zone (site 9) at
~23 km from sampling site 10, while the collar color cline
centers lie more than 50 km further along the transect near
the Río Changuinola (Figure 3F). The historical and contem-
porary CIs for beard length cline centers overlap, but the CIs
for cline widths do not, indicating that the cline has widened
by ~11 km (Supplementary Table S5). This widening could
indicate some degree of bidirectional introgression or relaxed
selection of beard length, but we do not currently have further
evidence to explain this change. Epaulet width is a shallow
cline that spans the whole hybrid zone and does not exhibit a
strong sigmoidal shape (Figure 3B). Cline estimates for epau-
let width were inconsistent across the analysis, with the best
cline model switching between model I and III depending on
the chain length and the addition of the contemporary sam-
pling site 9.5 (Supplementary Table S5). Given these results,
we cannot accurately resolve any temporal changes (or lack
thereof) in this trait.
In contrast, the olive belly coloration exhibits evidence
of movement across this hybrid zone over the last 30 years
(Figure 3E; Supplementary Table S5). The historical dataset
has an estimated cline center for belly color at 71.31 km
(95% CI: 67.74–75.47 km), while the contemporary dataset
has an estimated cline center at 80.47 km (95% CI: 77.89–
84.59 km). Based on nonoverlapping CIs, we conclude that
this cline has moved about 9 km in the last 30 years. The
cline for belly color also narrowed signicantly, with an esti-
mated historical width of 60.58 km (95% CI: 50.98–72.46
km) compared to only 8.73 km (95% CI: 1.55–17.02 km)
for the contemporary dataset. The difference in the contem-
porary and historical belly color clines was consistent with
whether or not sampling site 9.5 was included in the analy-
ses (Supplementary Table S5). Thus, the belly color cline has
shifted to the Río Changuinola (Figure 3F) and is now similar
in position and width to the collar color cline.
Discussion
The hybrid zone between the golden-collared manakin and
the white-collared manakin is one of the few documented
cases showing asymmetrical introgression of a male second-
ary sexual trait. In agreement with previous research on this
system (Brumeld et al., 2001; Parchman et al., 2013; Parsons
et al., 1993), we see that yellow collars, which have been
shown to be under positive sexual selection in this hybrid
zone (McDonald et al., 2001; Stein & Uy, 2006b, 2006a),
have been displaced up to the large regional river and dene
the phenotypic transition between the parental forms. We also
found evidence of the displacement of several genetic clines
toward this phenotypic transition, implying that there is his-
torical introgression across several independent loci through-
out the genome. Additionally, we found evidence that belly
plumage coloration, a male secondary sexual trait, has moved
and narrowed since the initial sampling, with belly color
changing to a darker green over time. Although the possibility
of direct sexual selection on belly color has not been formally
tested, dark bellies follow the same introgression pattern as
male yellow collar color. Together, these results suggest that
dark bellies are being selected in addition to yellow collars.
Despite this movement and displacement of putatively sex-
ually selected traits and loci, we found clear evidence that the
genomic transition of this hybrid zone and several phenotypic
clines have remained spatially stable over the span of ~25–30
years. The contrast between genomic stability and ongoing
introgression of a putatively sexually selected trait implies
that, while sexual selection is driving movement in several
traits and genetic loci towards the river, there is a counterac-
tive selective force maintaining the large-scale stability of the
hybrid zone. While the specic mechanisms maintaining the
stability of the hybrid zone are currently unknown, one coun-
teracting force could be generalized selection against hybrid
individuals through, for example, reduced reproductive suc-
cess (Svedin et al., 2008) and/or genetic incompatibilities
(Schumer et al., 2018). Independent of the specic mechanism
of selection, our work highlights how the interplay of sexual
selection and natural selection shape the genomes of hybrids
and the dynamics of the hybrid zone as a whole.
Asymmetrical introgression in avian tension zones
Previous studies characterizing the Manacus hybrid zone phe-
notypically and genetically found evidence of asymmetrical
introgression of male M. vitellinus plumage traits, yellow col-
lar color and olive belly color, into populations dominated
genetically by M. candei (Brumeld et al., 2001; Parchman
et al., 2013; Parsons et al., 1993). Despite this introgression,
we found that the Manacus hybrid zone remains narrow,
consistent with previous studies, which found that six out of
seven genetic markers had cline widths of less than 11 km
(Brumeld et al., 2001). Additionally, our genomic hybrid
index and admixture analysis corroborate the previous stud-
ies (Brumeld et al., 2001; Parchman et al., 2013; Parsons
et al., 1993), nding that sampling sites 4–8 are majority M.
candei-like, and that site 9 is the genomic hybrid center of the
hybrid zone. The results present a clear picture of very limited
genetic introgression across a geographically narrow, tempo-
rally stable hybrid zone.
Studies of other avian hybrid zones also report asym-
metrical introgression within narrow hybrid zones, e.g.,
northern ickers (Aguillon & Rohwer, 2022); red-backed
fairy-wrens (Baldassarre et al., 2014); long-tailed nches
(Grifth & Hooper, 2017); wagtails (Semenov et al., 2021);
jacanas (Lipshutz et al., 2019); and tanagers (Morales-Rozo
et al., 2017). Many of these hybrid zone studies posit that
movement may be controlled by changes in the environment
(Aguillon & Rohwer, 2022; Carling & Brumeld, 2008;
Carling & Zuckerberg, 2011; Walsh et al., 2020), or climate
change (Taylor et al., 2014); however, sexual selection appears
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
10 Long et al.
to play a more important role in Manacus. Our results better
align with other hybrid zones that show a trait under sex-
ual selection displaced from the genomic hybrid center due
to either intersexual (mate choice) or intrasexual (competi-
tion for access to mates) selection (Baldassarre et al., 2014;
Lipshutz et al., 2019; Yang et al., 2018). The Manacus hybrid
zone is a prime example of sexual selection driving key diag-
nostic traits across species boundaries; however, not all dis-
placed clines represent ongoing introgression (Semenov et al.,
2021). As such, reassessing hybrid zones with asymmetrical
introgression over multiple time points helps to clarify which
traits have ongoing introgression.
Barriers to gene flow despite positive selection
Male collar color displayed nearly identical clines in the con-
temporary and historical datasets; specically, a sharp tran-
sition in collar color from yellow to white occurred at the
Río Changuinola (Figure 1). Collar color likely introgressed
through positive sexual selection mediated by female choice
(Stein & Uy, 2006b), male aggression (McDonald et al.,
2001), or both. If positive sexual selection is still operating, as
it appears to be for belly color, collar color should continue to
spread. However, yellow collar plumage appears to be “stuck”
at the river, displaced by more than 50 km from the genomic
transition, consistent with previous studies (Brumeld et al.,
2001; Parchman et al., 2013; Parsons et al., 1993). Thus, the
Río Changuinola appears to present a limit to collar color
introgression, either by acting as a barrier to gene ow or
marking a transition in another relevant evolutionary factor
(Bennett et al., 2021). Belly color introgression may prove to
be similarly limited at the river.
Rivers have been shown as effective barriers to dispersal
in tropical birds (Haffer, 1997; Moncrieff et al., 2024; Naka
et al., 2012; Naka et al., 2022) and other taxa, e.g., frogs
(Fouquet et al., 2012), primates (Fordham et al., 2020), and
butteries (Rosser et al., 2021). However, the extent to which
a river, or any geographic barrier, serves to reduce gene ow
likely depends on the dispersal mode or capability of the
species of interest (Claramunt et al., 2012; Claramunt et al.,
2022; Naka et al., 2022; Nazareno et al., 2021). The golden-
collared manakin is capable of ying substantial distances
over open water (Moore et al., 2008), and we observed radio-
tagged females dispersing beyond a kilometer from the lek
where they were tagged. In one case, a female M. vitellinus
individual was tracked ying about 100 m over open cattle
pasture to a neighboring stand of trees to forage before ying
back to the lek in an isolated forest patch (KML personal
observation). Additionally, Manacus manakins are secondary
forest specialists, and most leks in our study area are along
waterways or anked by cacao plantations or cattle pastures.
Thus, while the Río Changuinola seems to be acting as a bar-
rier to gene ow given the location of the displaced clines,
Manacus manakins are capable of crossing the river, and there
have been a few cases recorded of yellow-collared males on
the west bank of the river (McDonald et al., 2001; Stein &
Uy, 2006b, KML personal observation). Therefore, dispersal
ability alone is insufcient to explain why clines under posi-
tive selection remain “stuck” at the Río Changuinola, instead
of spreading further beyond the river.
Besides dispersal constraints, other mechanisms have
been proposed as mediators for the stalling of introgres-
sion at the river. For example, Bennett et al. (2021) discuss
four hypotheses for the observed phenomenon, including
frequency-dependent selection resulting in differing favored
male phenotypes on each side of the river, geographic varia-
tion in the color or lighting conditions where males are per-
forming, differences in the male choice of display site location
for optimal display conspicuousness, and variation in female
preference on each side of the river. Presently, however, fur-
ther research is necessary to understand if any of the pro-
posed mechanisms, or a combination of them, may be playing
a role in stalling introgression at the Río Changuinola.
Melanization as a candidate for selection
Similar to the yellow collar in M. vitellinus, the belly color
introgressed from M. vitellinus (dark olive) into M. candei
(light yellow) and is displaced from the genomic center of
the hybrid zone (Brumeld et al., 2001). Here, we nd evi-
dence that belly color has continued to introgress, and its
cline center has moved ~9 km over the last 30 years, leading
to a current phenotypic transition at the Río Changuinola.
Given an estimated generation length of ~2.5 years, averaged
from estimates of M. candei and M. vitellinus from Bird et al.
(2020), this means that in about 12 generations, belly color
has introgressed about 0.7 km per generation. These ndings
suggest that the selection of belly color is ongoing in this
hybrid system.
Several nonexclusive factors could explain the selection for
dark olive bellies in hybrid populations from natural or sex-
ual selection. For natural selection, there could be an ecologi-
cal or environmental advantage to having darker olive bellies.
For sexual selection, both female choice and male–male com-
petition have been implicated as important in lek-breeding
systems and the Manacus system in particular (McDonald
et al., 2001; Stein & Uy, 2006a, 2006b). We propose a few
hypotheses for sexual selection: (1) there is a sexual advan-
tage of a yellow collar that is enhanced by contrast with a
dark belly, so both introgress together, (2) females prefer dark
olive bellies over yellow bellies independently from collar
plumage, (3) males with olive bellies are more aggressive and
could benet from male–male competition. We expand on the
male–male aggression/melanization hypothesis below.
The olive color in Manacus dark bellies is proposed to
result from melanin deposition on yellow feathers (Bennett et
al., 2021). Pathways controlling aggression and melanin pro-
duction share some of the same upstream regulators (Ducrest
et al., 2008), which has been proposed as an explanation for
the observed pattern that animals with darker colorations
tend to be more aggressive, especially in birds (Da Silva et al.,
2013; de Zwaan et al., 2019; Ducrest et al., 2008; Mai et al.,
2011; Mateos-Gonzalez & Senar, 2012). Whether manakins
with darker bellies are more aggressive remains to be tested
directly, but two previous studies performed aggression assays
on Manacus manakins based on hybrid status. McDonald et
al. (2001) found that hybrid males from the region of yellow
plumage introgression were the most aggressive to the presen-
tation of a mount, followed by M. vitellinus males and, lastly,
M. candei males. Stein and Uy (2006b) performed a similar
experiment and found no difference in aggression between
yellow and white-collared males; however, they worked in a
different location along the upper Río Changuinola where
the river is narrow and mixed leks with yellow and white-
collared birds are found.
If aggression in Manacus is inuenced by melanization
(or vice versa), then the difference in response of the manak-
ins tested in the two previous studies could be explained by
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
Evolution (2024), Vol. XX 11
differences in the introgression of dark melanized bellies.
Specically, McDonald et al. (2001) tested yellow-collared
males near sampling site 5, where belly color is typically
darker than in leks along the upper Río Changuinola where
Stein and Uy (2006b) worked. Both studies focused on
yellow-collared hybrids, but neither quantied the belly color
of the birds tested. If aggression is linked to green bellies/
increased melanization more so than yellow collars, their
results may be reconcilable. Careful experiments that directly
test female preference for olive bellies, male aggression tri-
als with varying olive belly colors, as well as categorizing the
genes and metabolic pathways responsible for melanization
in this hybrid zone and its causal (or lack thereof) link to
aggression will be necessary to tease apart the connection of
introgressing olive bellies, melanization, male–male aggres-
sion, and the subsequent effect on hybrid zone dynamics.
Independent of the evolutionary force underlying the move-
ment of belly coloration across the hybrid zone, it is notable
that belly color has appeared to have lagged behind collar
color. There are several possible explanations for this dispa-
rate shift in their movement, including changes in selective
pressures (e.g., female choice) over time and/or differences in
the genomic architecture of these two coloration traits. First,
the traits that females nd attractive could change over time
or vary within populations (Chaine & Lyon, 2008; DuVal et
al., 2023). Within the context of the Manacus system, belly
color could be under recent selection, while collar color was
under stronger sexual selection historically. Another expla-
nation for the lagging introgression of belly color could be
due to differences in the genomic architecture between the
two traits. For example, the two traits could be the product
of independent genomic pathways, as seen in melanin and
carotenoid processing genes in Setophaga warblers (Baiz et
al., 2021). If the loci controlling belly color and collar color
are under weak linkage, they would appear to behave as inde-
pendent units in the genome and thus move at different rates
if under differing selective pressures. Additionally, the genetic
architecture underlying both of these traits in Manacus is
presently unknown, which may inuence the movement of a
phenotype if the trait is highly polygenic (Sachdeva & Barton,
2018). Interestingly, we observed that belly color has higher
variation than collar color in Manacus, a pattern also noted
by Brumeld et al. (2001). At sites near the Río Changuinola,
male belly color exhibits greater within-site variation than
collar color, which may indicate that belly color is in fact,
polygenic in this system. The underlying genomic archi-
tecture, along with female choice, has putatively impacted
these traits, but it is not yet fully understood and could yield
important insights into the dynamics of hybrid systems under
sexual selection.
Evolutionary forces and hybrid zone dynamics
The Manacus hybrid zone is an ideal system to study the
effects of sexual selection, natural selection, and the resulting
asymmetrical introgression in a naturally occurring hybrid
zone. In addition to emerging genomic resources, this hybrid
zone benets from multiyear sampling, which is the most reli-
able way to empirically assess hybrid zone movement (Buggs,
2007). Our results indicate that there is an interplay between
sexual selection, which leads to the decoupling and move-
ment of certain traits and loci, and other evolutionary forces
(such as generalized selection against hybrids), which keep the
majority of the hybrid genome in relative stasis.
The stability (or lack thereof) of hybrid zones is likely the
product of conicting evolutionary forces. Selection of some
traits can be counteracted by generalized selection against
hybrids, which in turn produces differential tness among
hybrid individuals and directly affects the movement of
alleles and traits across the landscape (Buggs, 2007). Sexual
selection is one mechanism that can lead to differential repro-
ductive tness in hybrids. Sexual selection, for example, can
reduce hybrid mating success by creating disfavored combi-
nations of traits (Naisbit et al., 2001; Svedin et al., 2008) or
by reducing the competitiveness of male sperm (Howard et
al., 1998). Yet, despite selection against hybrids being a com-
mon outcome of hybridization, sexual selection in hybrid
zones can simultaneously fuel the introgression and move-
ment of certain alleles, imparting adaptive potential to new
populations (Enard & Petrov, 2018; Pardo-Diaz et al., 2012;
Walsh et al., 2018). Overall, hybrid zones serve as a powerful
evolutionary model to study how similar evolutionary forces
can have a myriad of outcomes with heterogeneous responses
across different parts of the genome. Our work here in the
Manacus hybrid zone demonstrates that traits under sexual
selection can move over space and time despite stability in
the rest of the genome.
Hybrid zones are ideal cases to observe the dynamics
of evolution in action, but evolutionary forces can change
over time. Perceived outcomes may shift depending on the
population sampled and/or the time at which they were
sampled, highlighting the importance of following the tra-
jectories of hybrid populations as they evolve (Enbody et
al., 2023; Schumer et al., 2017). In this study, we com-
bined sampling over multiple time points and across geo-
graphic space, along with both phenotypic and genomic
data, to shed light on speciation and selection dynamics
in a wild bird. Our results reveal an uncoupling of sexual
and natural selection in a natural hybrid zone and evidence
for ongoing introgression of dark belly plumage. Despite
genomic stability in its center, the Manacus hybrid zone is
a dynamic system with phenotypes that continue to change
over a short timescale. We believe longitudinal sampling
in other systems will reveal similar evolutionary dynam-
ics and facilitate a deeper understanding of how evolution
works in nature.
Supplementary material
Supplementary material is available online at Evolution.
Data availability
DNA sequence data les are available under the NCBI
BioProject PRJNA893627 and PRJNA951544. Scripts for
analyses and gures are available at https://github.com/
kiralong/manacus_hz_movement_ms.
Author contributions
K.M.L., M.J.B., and J.D.B. conceived and designed the study.
K.M.L. collected eld data, made graphics, and wrote the
original draft of the manuscript. K.F.P.B. contributed lab
work and sequencing data. K.M.L. and A.G.R.C. performed
lab work and analyzed the data. A.G.R.C., K.F.P.B., J.M.C.,
M.J.B., and J.D.B. provided discussions and feedback on the
text and analyses. All authors approved the nal manuscript.
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
12 Long et al.
Funding
This project was supported by the NSF DGE 10-69157
IGERT to K.M.L. and A.G.R.C., the USDA National Institute
of Food and Agriculture, Hatch project 1026333 (ILLU-875-
984) to J.D.B., the Smithsonian Institution and a National
Museum of Natural History Core grant to M.J.B.
Conict of interest: The authors declare no conict of interest.
Acknowledgments
We thank the Smithsonian Tropical Research Institute,
including Owen McMillan, Lil Camacho, Raineldo Urriola,
and Isis Ochoa, for logistics and MiAmbiente for permit-
ting. We thank Cesar Romero for logistics and hospitality,
and our eld assistants: Adolfo Muñoz, Ovidio Jaramillo,
André Nguyen, Alexander Worm, Ivy Ciaburri, Olivia Ferrari,
Elena Prado-Ragan, Vaughn Hage, Luis Ramos Vazquez,
and Evalynn Trumbo. We are grateful to the local families
and communities who supported our work, including but
not limited to the Jimenez family, the Lezcano family, the
Aguilar family, the López family, the Teribe people, and the
Ngäbe-Buglé people. Thank you to the staff at the University
of Illinois at Urbana-Champaign departments of Natural
Resources and Environmental Sciences, Integrative Biology,
and the Smithsonian’s National Museum of Natural History,
including Mary Faith Flores and Carrie Craig. Additionally,
we thank Becky Fuller, Troy Murphy, Tom Parsons, Peri
Bolton, and Karthik Yarlagadda for helpful discussions and
advice on this manuscript.
References
Abbott, R. J., Albach, D., Ansell, S., Arntzen, J. W., Baird, S. J. E., Bierne,
N., & Zinner, D. (2013). Hybridization and speciation. Journal of
Evolutionary Biology, 26(2), 229–246. https://doi.org/10.1111/
j.1420-9101.2012.02599.x
Abbott, R. J., Barton, N. H., & Good, J. M. (2016). Genomics of
hybridization and its evolutionary consequences. Molecular Ecol-
ogy, 25(11), 2325–2332. https://doi.org/10.1111/mec.13685
Aguillon, S. M., & Rohwer, V. G. (2022). Revisiting a classic hybrid zone:
Movement of the northern icker hybrid zone in contemporary times.
Evolution, 76(5), 1082–1090. https://doi.org/10.1111/evo.14474
Alexander, D. H., Novembre, J., & Lange, K. (2009). Fast model-based
estimation of ancestry in unrelated individuals. Genome Research,
19(9), 1655–1664. https://doi.org/10.1101/gr.094052.109
Bailey, R. I. (2020). gghybrid: R package for evolutionary analysis of
hybrids and hybrid zones. https://doi.org/10.5281/zenodo.3676499
Baird, N. A., Etter, P. D., Atwood, T. S., Currey, M. C., Shiver, A. L.,
Lewis, Z. A., Johnson, E. A., Cresko, W. A., & Johnson, E. A.
(2008). Rapid SNP discovery and genetic mapping using sequenced
RAD markers. PLoS One, 3(10), e3376. https://doi.org/10.1371/
journal.pone.0003376
Baiz, M. D., Wood, A. W., Brelsford, A., Lovette, I. J., & Toews, D.
P. L. (2021). Pigmentation genes show evidence of repeated diver-
gence and multiple bouts of introgression in Setophaga warblers.
Current Biology: CB, 31(3), 643–649.e3. https://doi.org/10.1016/j.
cub.2020.10.094
Baldassarre, D. T., White, T. A., Karubian, J., & Webster, M. S. (2014).
Genomic and morphological analysis of a semipermeable avian
hybrid zone suggests asymmetrical introgression of a sexual signal.
Evolution, 68(9), 2644–2657. https://doi.org/10.1111/evo.12457
Barton, N. H. (1979). The dynamics of hybrid zones. Heredity, 43(3),
341–359. https://doi.org/10.1038/hdy.1979.87
Barton, N. H., & Hewitt, G. M. (1985). Analysis of hybrid zones.
Annual Review of Ecology, Evolution, and Systematics, 16, 113–
148. www.annualreviews.org
Bennett, K. F. P., Lim, H. C., & Braun, M. J. (2021). Sexual selection
and introgression in avian hybrid zones: Spotlight on Manacus.
Integrative and Comparative Biology, 61(4), 1291–1309. https://
doi.org/10.1093/icb/icab135
Billerman, S. M., Cicero, C., Bowie, R. C. K., & Carling, M. D. (2019).
Phenotypic and genetic introgression across a moving woodpecker
hybrid zone. Molecular Ecology, 28(7), 1692–1708. https://doi.
org/10.1111/mec.15043
Bird, J. P., Martin, R., Akçakaya, H. R., Gilroy, J., Bureld, I. J., Garnett,
S. T., Butchart, S. H. M., Taylor, J., Şekercioğlu, C. H., & Butchart,
S. H. M. (2020). Generation lengths of the world’s birds and their
implications for extinction risk. Conservation Biology: The Journal
of the Society for Conservation Biology, 34(5), 1252–1261. https://
doi.org/10.1111/cobi.13486
Brumeld, R. T., Jernigan, R. W., McDonald, D. B., & Braun, M. J.
(2001). Evolutionary implications of divergent clines in an avian
(Manacus: Aves) hybrid zone. Evolution, 55(10), 2070–2087.
https://doi.org/10.1111/j.0014-3820.2001.tb01322.x
Buggs, R. J. A. (2007). Empirical study of hybrid zone movement.
Heredity, 99(3), 301–312. https://doi.org/10.1038/sj.hdy.6800997
Carling, M. D., & Brumeld, R. T. (2008). Haldane’s rule in an avian
system: Using cline theory and divergence population genetics to
test for differential introgression of mitochondrial, autosomal,
and sex-linked loci across the Passerina bunting hybrid zone.
Evolution, 62(10), 2600–2615. https://doi.org/10.1111/j.1558-
5646.2008.00477.x
Carling, M. D., & Zuckerberg, B. (2011). Spatio-temporal changes in
the genetic structure of the Passerina bunting hybrid zone. Molec-
ular Ecology, 20(6), 1166–1175. https://doi.org/10.1111/j.1365-
294X.2010.04987.x
Chaine, A. S., & Lyon, B. E. (2008). Adaptive plasticity in female mate
choice dampens sexual selection on male ornaments in the lark
bunting. Science, 319(5862), 459–462. https://doi.org/10.1126/
science.1149167
Chan, K. O., Hutter, C. R., Wood, P. L., Grismer, L. L., Das, I., & Brown,
R. M. (2020). Gene ow creates a mirage of cryptic species in a
Southeast Asian spotted stream frog complex. Molecular Ecology,
29(20), 3970–3987. https://doi.org/10.1111/mec.15603
Chang, C. C., Chow, C. C., Tellier, L. C. A. M., Vattikuti, S., Purcell, S.
M., & Lee, J. J. (2015). Second-generation PLINK: Rising to the
challenge of larger and richer datasets. GigaScience, 4(1), 1–16.
https://doi.org/10.1186/S13742-015-0047-8
Claramunt, S., Derryberry, E. P., Remsen, J. V., & Brumeld, R. T.
(2012). High dispersal ability inhibits speciation in a continen-
tal radiation of passerine birds. Proceedings of the Royal Soci-
ety B: Biological Sciences, 279(1733), 1567–1574. https://doi.
org/10.1098/rspb.2011.1922
Claramunt, S., Hong, M., & Bravo, A. (2022). The effect of ight ef-
ciency on gapcrossing ability in Amazonian forest birds. Biotrop-
ica, 54(4), 860–868. https://doi.org/10.1111/btp.13109
Cook, E. G., Murphy, T. G., & Johnson, M. A. (2013). Colorful displays
signal male quality in a tropical anole lizard. Naturwissenschaften,
100(10), 993–996. https://doi.org/10.1007/s00114-013-1095-5
Coyne, J. A., & Orr, H. A. (2004). Speciation. Sinauer Associates.
Da Silva, A., van den Brink, V., Emaresi, G., Luzio, E., Bize, P., Dreiss,
A. N., & Roulin, A. (2013). Melanin-based colour polymorphism
signals aggressive personality in nest and territory defence in the
tawny owl (Strix aluco). Behavioral Ecology and Sociobiology,
67(7), 1041–1052. https://doi.org/10.1007/s00265-013-1529-2
Dasmahapatra, K. K., Blum, M. J., Aiello, A., Hackwell, S., Davies,
N., Bermingham, E. P., & Mallet, J. (2002). Inferences from a rap-
idly moving hybrid zone. Evolution, 56(4), 741–753. https://doi.
org/10.1554/0014-3820(2002)056[0741:ifarmh]2.0.co;2
Day, L. B., Helmhout, W., Pano, G., Olsson, U., Hoeksema, J. D., &
Lindsay, W. R. (2021). Correlated evolution of acrobatic display
and both neural and somatic phenotypic traits in Manakins (Pip-
ridae). Integrative and Comparative Biology, 61(4), 1343–1362.
https://doi.org/10.1093/icb/icab139
de Zwaan, D. R., Barnes, S., & Martin, K. (2019). Plumage melanism is
linked to male quality, female parental investment and assortative
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
Evolution (2024), Vol. XX 13
mating in an alpine songbird. Animal Behaviour, 156, 41–49.
https://doi.org/10.1016/j.anbehav.2019.06.034
Derryberry, E. P., Derryberry, G. E., Maley, J. M., & Brumeld, R.
T. (2014). hzar: Hybrid zone analysis using an R software pack-
age. Molecular Ecology Resources, 14(3), 652–663. https://doi.
org/10.1111/1755-0998.12209
Doucet, S. M., & Hill, G.E. (2009). Do museum specimens accurately
represent wild birds? A case study of carotenoid, melanin, and
structural colours in long-tailed manakins Chiroxiphia linearis.
Journal of Avian Biology, 40, 146–156. https://doi.org/10.1111/
j.1600-048X.2009.03763.x
Ducrest, A., Keller, L., & Roulin, A. (2008). Pleiotropy in the mela-
nocortin system, coloration and behavioural syndromes. Trends in
Ecology & Evolution, 23(9), 502–510. https://doi.org/10.1016/j.
tree.2008.06.001
DuVal, E. H., Fitzpatrick, C. L., Hobson, E. A., & Servedio, M. R.
(2023). Inferred attractiveness: A generalized mechanism for sexual
selection that can maintain variation in traits and preferences over
time. PLoS Biology, 21(10), e3002269. https://doi.org/10.1371/
journal.pbio.3002269
Enard, D., & Petrov, D. A. (2018). Evidence that RNA viruses drove
adaptive introgression between Neanderthals and modern
humans. Cell, 175(2), 360–371.e13. https://doi.org/10.1016/j.
cell.2018.08.034
Enbody, E. D., Sendell-Price, A. T., Sprehn, C. G., Rubin, C. J.,
Visscher, P. M., Grant, B. R., Grant, P. R., & Andersson, L. (2023).
Community- wide genome sequencing reveals 30 years of Dar-
win’s nch evolution. Science, 381(6665), eadf6218. https://doi.
org/10.1126/science.adf6218
Engebretsen, K. N., Barrow, L. N., Rittmeyer, E. N., Brown, J. M., &
Lemmon, E. M. (2016). Quantifying the spatiotemporal dynamics
in a chorus frog (Pseudacris) hybrid zone over 30 years. Ecology and
Evolution, 6(14), 5013–5031. https://doi.org/10.1002/ece3.2232
Etter, P. D., Preston, J. L., Bassham, S., Cresko, W. A., & Johnson, E.
A. (2011). Local de novo assembly of RAD paired-end contigs
using short sequencing reads. PLoS One, 6(4), e18561. https://doi.
org/10.1371/journal.pone.0018561
Fitzpatrick, B. M. (2012). Estimating ancestry and heterozygosity of
hybrids using molecular markers. BMC Evolutionary Biology,
12(1), 131. https://doi.org/10.1186/1471-2148-12-131
Fordham, G., Shanee, S., & Peck, M. (2020). Effect of river size on
Amazonian primate community structure: A biogeographic analy-
sis using updated taxonomic assessments. American Journal of Pri-
matology, 82(7), e23136. https://doi.org/10.1002/ajp.23136
Fouquet, A., Ledoux, J. -B., Dubut, V., Noonan, B. P., & Scotti, I.
(2012). The interplay of dispersal limitation, rivers, and historical
events shapes the genetic structure of an Amazonian frog. Biolog-
ical Journal of the Linnean Society, 106(2), 356–373. https://doi.
org/10.1111/j.1095-8312.2012.01871.x
Fridolfsson, A.-K., & Ellegren, H. (1999). A simple and universal
method for molecular sexing of non-ratite birds. Journal of Avian
Biology, 30(1), 116. https://doi.org/10.2307/3677252
Gompert, Z., & Buerkle, A. (2010). Introgress: A software package
for mapping components of isolation in hybrids. Molecular Ecol-
ogy Resources, 10(2), 378–384. https://doi.org/10.1111/j.1755-
0998.2009.02733.x
Grifth, S. C., & Hooper, D. M. (2017). Geographical variation in bill
colour in the Long-tailed Finch: Evidence for a narrow zone of
admixture between sub-species. Emu - Austral Ornithology, 117(2),
141–150. https://doi.org/10.1080/01584197.2016.1277763
Gross, B. L., & Rieseberg, L. H. (2005). The ecological genetics of
Homoploid hybrid speciation. The Journal of Heredity, 96(3),
241–252. https://doi.org/10.1093/jhered/esi026
Haffer, J. (1997). Contact zones between birds of Southern Ama-
zonia. Ornithological Monographs, 48, 281–305. https://doi.
org/10.2307/40157539
Harr, B., & Price, T. (2014). Climate change: A hybrid zone moves
north. Current Biology: CB, 24(6), R230–R232. https://doi.
org/10.1016/j.cub.2014.02.023
Hohenlohe, P. A., Bassham, S., Currey, M., & Cresko, W. A. (2012).
Extensive linkage disequilibrium and parallel adaptive divergence
across threespine stickleback genomes. Philosophical Transactions
of the Royal Society of London, Series B: Biological Sciences,
367(1587), 395–408. https://doi.org/10.1098/rstb.2011.0245
Howard, D. J., Gregory, P. G., Chu, J., & Cain, M. L. (1998). Con-
specic sperm precedence is an effective barrier to hybridization
between closely related species. Evolution, 52(2), 511–516. https://
doi.org/10.1111/j.1558-5646.1998.tb01650.x
Jombart, T. (2008). adegenet: A R package for the multivariate analysis
of genetic markers. Bioinformatics, 24(11), 1403–1405. https://doi.
org/10.1093/bioinformatics/btn129
Jombart, T., & Ahmed, I. (2011). adegenet 1.3-1: New tools for the
analysis of genome-wide SNP data. Bioinformatics, 27(21), 3070–
3071. https://doi.org/10.1093/bioinformatics/btr521
Kearns, A. M., Restani, M., Szabo, I., Schrøder-Nielsen, A., Kim, J. A.,
Richardson, H. M., Omland, K. E., Fleischer, R. C., Johnsen, A.,
& Omland, K. E. (2018). Genomic evidence of speciation rever-
sal in ravens. Nature Communications, 9(1), 906. https://doi.
org/10.1038/s41467-018-03294-w
Kelly, R. J., Murphy, T. G., Tarvin, K. A., & Burness, G. (2012).
Carotenoid- based ornaments of female and male American gold-
nches (Spinus tristis) show sex-specic correlations with immune
function and metabolic rate. Physiological and Biochemical Zool-
ogy: PBZ, 85(4), 348–363. https://doi.org/10.1086/666059
Kirwan, G. M., & Green, G. (2011). Cotingas and Manakins. Princeton
University Press.
Lavretsky, P., Janzen, T., & McCracken, K. G. (2019). Identifying
hybrids & the genomics of hybridization: Mallards & American
black ducks of Eastern North America. Ecology and Evolution,
9(6), 3470–3490. https://doi.org/10.1002/ece3.4981
Lavretsky, P., Peters, J. L., Winker, K., Bahn, V., Kulikova, I., Zhuravlev, Y.
N., McCracken, K. G., Barger, C., Gurney, K., & McCracken, K. G.
(2016). Becoming pure: Identifying generational classes of admixed
individuals within lesser and greater scaup populations. Molecular
Ecology, 25(3), 661–674. https://doi.org/10.1111/mec.13487
Li, H. (2013). Aligning sequence reads, clone sequences and assem-
bly contigs with BWA-MEM. ArXiv preprint: not peer reviewed.
https://doi.org/10.48550/arXiv.1303.3997
Li, H., & Durbin, R. (2009). Fast and accurate short read alignment
with Burrows-Wheeler transform. Bioinformatics, 25(14), 1754–
1760. https://doi.org/10.1093/bioinformatics/btp324
Li, H., Handsaker, B., Wysoker, A., Fennell, T., Ruan, J., Homer, N.,
Durbin, R., Abecasis, G., & Durbin, R.; 1000 Genome Project Data
Processing Subgroup. (2009). The sequence alignment/map format
and SAMtools. Bioinformatics, 25(16), 2078–2079. https://doi.
org/10.1093/bioinformatics/btp352
Lipshutz, S. E., Meier, J. I., Derryberry, G. E., Miller, M. J., Seehau-
sen, O., & Derryberry, E. P. (2019). Differential introgression of a
female competitive trait in a hybrid zone between sexrole reversed
species. Evolution, 73(2), 188–201. https://doi.org/10.1111/
evo.13675
Longmire, J. L., Maltbie, M., & Baker, R. J. (1997). Use of “lysis buf-
fer” in isolation and its implications for museum collections. Occa-
sional papers of the Museum of Texas Tech University. (Vol. 163).
Museum of Texas Tech University.
Macholán, M., Baird, S. J., Munclinger, P., Dufková, P., Bímová, B., &
Piálek, J. (2008). Genetic conict outweighs heterogametic incom-
patibility in the mouse hybrid zone? BMC Evolutionary Biology,
8(1), 271. https://doi.org/10.1186/1471-2148-8-271
Macholán, M., Munclinger, P., Šugerková, M., Dufková, P., Bímová,
B., Božíková, E., Piálek, J., & Piálek, J. (2007). Genetic analysis
of autosomal and x-linked markers across a mouse hybrid zone.
Evolution, 61(4), 746–771. https://doi.org/10.1111/j.1558-
5646.2007.00065.x
Mai, A., Wakamatsu, K., & Roulin, A. (2011). Melanin-based color-
ation predicts aggressiveness and boldness in captive eastern Her-
mann’s tortoises. Animal Behaviour, 81(4), 859–863. https://doi.
org/10.1016/j.anbehav.2011.01.025
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
14 Long et al.
Mateos-Gonzalez, F., & Senar, J. C. (2012). Melanin-based trait pre-
dicts individual exploratory behaviour in siskins, Carduelis spinus.
Animal Behaviour, 83(1), 229–232. https://doi.org/10.1016/j.anbe-
hav.2011.10.030
McDonald, D. B., Clay, R. P., Brumeld, R. T., & Braun, M. J. (2001).
Sexual selection on plumage and behavior in an avian hybrid zone:
Experimental tests of male-male interactions. Evolution, 55(7),
1443–1451. https://doi.org/10.1111/j.0014-3820.2001.tb00664.x
Metzler, D., Knief, U., Peñalba, J. V., & Wolf, J. B. W. (2021). Assorta-
tive mating and epistatic matingtrait architecture induce complex
movement of the crow hybrid zone. Evolution, 75(12), 3154–3174.
https://doi.org/10.1111/evo.14386
Moncrieff, A. E., Faircloth, B. C., Remsen, R. C., Hiller, A. E., Felix,
C., Capparella, A. P., Brumeld, R. T., Valqui, T., & Brumeld, R.
T. (2024). Implications of headwater contact zones for the river-
ine barrier hypothesis: A case study of the Blue-capped Manakin
(Lepidothrix coronata). Evolution, 78(1), 53–68. https://doi.
org/10.1093/evolut/qpad187
Moore, R. P., Robinson, W. D., Lovette, I. J., & Robinson, T. R. (2008).
Experimental evidence for extreme dispersal limitation in trop-
ical forest birds. Ecology Letters, 11(9), 960–968. https://doi.
org/10.1111/j.1461-0248.2008.01196.x
Morales-Rozo, A., Tenorio, E. A., Carling, M. D., & Cadena, C. D.
(2017). Origin and cross-century dynamics of an avian hybrid zone.
BMC Evolutionary Biology, 17(1), 257. https://doi.org/10.1186/
s12862-017-1096-7
Moran, B. M., Payne, C., Langdon, Q., Powell, D. L., Brandvain, Y., &
Schumer, M. (2021). The genomic consequences of hybridization.
ELife, 10, 10. https://doi.org/10.7554/eLife.69016
Naisbit, R. E., Jiggins, C. D., & Mallet, J. (2001). Disruptive sexual
selection against hybrids contributes to speciation between Helico-
nius cydno and Heliconius melpomene. Proceedings of the Royal
Society B: Biological Sciences, 268(1478), 1849–1854. https://doi.
org/10.1098/rspb.2001.1753
Naka, L. N., Bechtoldt, C. L., Henriques, L. M. P., & Brumeld, R. T.
(2012). The role of physical barriers in the location of avian suture
zones in the Guiana Shield, Northern Amazonia. The American
Naturalist, 179(4), E115–E132. https://doi.org/10.1086/664627
Naka, L. N., Costa, B. M. S., Lima, G. R., & Claramunt, S. (2022). Riv-
erine barriers as obstacles to dispersal in Amazonian birds. Fron-
tiers in Ecology and Evolution, 10, 539. https://doi.org/10.3389/
fevo.2022.846975
Nazareno, A. G., Knowles, L. L., Dick, C. W., & Lohmann, L. G. (2021).
By animal, water, or wind: Can dispersal mode predict genetic con-
nectivity in riverine plant species? Frontiers in Plant Science, 12, 31.
https://doi.org/10.3389/fpls.2021.626405
Ottenburghs, J., Kraus, R. H. S., van Hooft, P., van Wieren, S. E., Yden-
berg, R. C., & Prins, H. H. T. (2017). Avian introgression in the
genomic era. Avian Research, 8(1), 30. https://doi.org/10.1186/
s40657-017-0088-z
Parchman, T. L., Gompert, Z., Braun, M. J., Brumeld, R. T., McDon-
ald, D. B., Uy, J. A. C., Buerkle, C. A., Jarvis, E. D., Schlinger, B. A.,
& Buerkle, C. A. (2013). The genomic consequences of adaptive
divergence and reproductive isolation between species of manakins.
Molecular Ecology, 22(12), 3304–3317. https://doi.org/10.1111/
mec.12201
Pardo-Diaz, C., Salazar, C., Baxter, S. W., Merot, C., Figueiredo-Ready,
W., Joron, M., Jiggins, C. D., & Jiggins, C. D. (2012). Adaptive
introgression across species boundaries in Heliconius butteries.
PLoS Genetics, 8(6), e1002752. https://doi.org/10.1371/journal.
pgen.1002752
Parsons, T. J., Olson, S. L., & Braun, M. J. (1993). Unidirectional
spread of secondary sexual plumage traits across an avian hybrid
zone. Science, 260(5114), 1643–1646. https://doi.org/10.1126/sci-
ence.260.5114.1643
Payseur, B. A., & Rieseberg, L. H. (2016). A genomic perspective on
hybridization and speciation. Molecular Ecology, 25(11), 2337–
2360. https://doi.org/10.1111/mec.13557
Phillips, B. L., Baird, S. J. E., & Moritz, C. (2004). When vicars meet:
A narrow contact zone between morphologically cryptic phylogeo-
graphic lineages of the rainforest skink, Carlia rubrigularis. Evolu-
tion, 58(7), 1536–1548. https://doi.org/10.1554/02-498
Pinto, B. J., Titus-McQuillan, J., Daza, J. D., & Gamble, T. (2019).
Persistence of a geographically-stable hybrid zone in Puerto Rican
dwarf geckos. The Journal of Heredity, 110(5), 523–534. https://
doi.org/10.1093/jhered/esz015
R Core Team. (2022). R: A language and environment for statisti-
cal computing. R Foundation for Statistical Computing. https://
www.r-project.org/
Reudink, M. W., Mech, S. G., Mullen, S. P., & Curry, R. L. (2007).
Structure and dynamics of the hybrid zone between black-capped
chickadee (Poecile atricapillus) and Carolina chickadee (P. carolin-
ensis) in Southeastern Pennsylvania. The Auk, 124(2), 463–478.
https://doi.org/10.1642/0004-8038(2007)124[463:sadoth]2.0.co;2
Roberts, N. S., & Mendelson, T. C. (2020). Reinforcement in the
banded darter Etheostoma zonale: The effect of sex and sympatry
on preferences. Ecology and Evolution, 10(5), 2499–2512. https://
doi.org/10.1002/ece3.6076
Rochette, N. C., RiveraColón, A. G., & Catchen, J. M. (2019). Stacks
2: Analytical methods for pairedend sequencing improve RAD-
seqbased population genomics. Molecular Ecology, 28(21), 4737–
4754. https://doi.org/10.1111/mec.15253
Rosser, N., Dasmahapatra, K. K., & Mallet, J. (2014). Stable Helico-
nius buttery hybrid zones are correlated with a local rainfall peak
at the edge of the Amazon basin. Evolution, 68(12), 3470–3484.
https://doi.org/10.1111/evo.12539
Rosser, N., Shirai, L. T., Dasmahapatra, K. K., Mallet, J., & Freitas,
A. V. L. (2021). The Amazon river is a suture zone for a polyphy-
letic group of co-mimetic heliconiine butteries. Ecography, 44(2),
177–187. https://doi.org/10.1111/ECOG.05282
Ruegg, K. (2008). Genetic, morphological, and ecological characteriza-
tion of a hybrid zone that spans a migratory divide. Evolution, 62(2),
452–466. https://doi.org/10.1111/j.1558-5646.2007.00263.x
Sachdeva, H., & Barton, N. H. (2018). Replicability of introgression
under linked, polygenic selection. Genetics, 210(4), 1411–1427.
https://doi.org/10.1534/genetics.118.301429
Schumer, M., Powell, D. L., Delclós, P. J., Squire, M., Cui, R., Andol-
fatto, P., & Rosenthal, G. G. (2017). Assortative mating and
persistent reproductive isolation in hybrids. Proceedings of the
National Academy of Sciences, 114(41), 10936–10941. https://doi.
org/10.1073/pnas.1711238114
Schumer, M., Xu, C., Powell, D. L., Durvasula, A., Skov, L., Holland,
C., Przeworski, M., Sankararaman, S., Andolfatto, P., Rosenthal,
G. G., & Przeworski, M. (2018). Natural selection interacts with
recombination to shape the evolution of hybrid genomes. Science,
360(6389), 656–660. https://doi.org/10.1126/science.aar3684
Seehausen, O., Alphen, J. J. M. van, & Witte, F. (1997). Cichlid sh
diversity threatened by eutrophication that curbs sexual selec-
tion. Science, 277(5333), 1808–1811. https://doi.org/10.1126/sci-
ence.277.5333.1808
Semenov, G. A., Linck, E., Enbody, E. D., Harris, R. B., Khaydarov, D.
R., Alström, P., Taylor, S. A., & Taylor, S. A. (2021). Asymmetric
introgression reveals the genetic architecture of a plumage trait.
Nature Communications, 12(1), 1019. https://doi.org/10.1038/
s41467-021-21340-y
Smith, K. L., Hale, J. M., Kearney, M. R., Austin, J. J., & Melville,
J. (2013). Molecular patterns of introgression in a classic hybrid
zone between the Australian tree frogs, Litoria ewingii and L. par-
aewingi: Evidence of a tension zone. Molecular Ecology, 22(7),
1869–1883. https://doi.org/10.1111/mec.12176
Snow, D. (2004). Family Pipridae (Manakins). In J. del Hoyo, A.
Elliott, & D. A. Christie (Eds.), Handbook of the birds of the
world. Vol. 9. Cotingas to pipits and wagtails (pp. 110–169). Lynx
Edicions.
Stein, A. C., & Uy, J. A. C. (2006a). Plumage brightness predicts male
mating success in the lekking golden-collared Manakin, Mana-
cus vitellinus. Behavioral Ecology, 17(1), 41–47. https://doi.
org/10.1093/beheco/ari095
Stein, A. C., & Uy, J. A. C. (2006b). Unidirectional introgression of a
sexually selected trait across an avian hybrid zone: A role for female
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
Evolution (2024), Vol. XX 15
choice? Evolution, 60(7), 1476–1485. https://doi.org/10.1554/05-
575.1
Svedin, N., Wiley, C., Veen, T., Gustafsson, L., & Qvarnström, A.
(2008). Natural and sexual selection against hybrid ycatchers.
Proceedings of the Royal Society B: Biological Sciences, 275(1635),
735–744. https://doi.org/10.1098/rspb.2007.0967
Szymura, J. M., & Barton, N. H. (1986). Genetic analysis of a hybrid
zone between the re-bellied toads, Bombina bombina and B. var-
iegata, near Cracow in Southern Poland. Evolution, 40(6), 1141–
1159. https://doi.org/10.1111/j.1558-5646.1986.tb05740.x
Szymura, J. M., & Barton, N. H. (1991). The genetic structure of the
hybrid zone between the rebellied toads Bombina bombina and B.
variegata: Comparisons between transects and between loci. Evo-
lution, 45(2), 237–261. https://doi.org/10.1111/j.1558-5646.1991.
tb04400.x
Taylor, S. A., & Larson, E. L. (2019). Insights from genomes into
the evolutionary importance and prevalence of hybridization in
nature. Nature Ecology & Evolution, 3(2), 170–177. https://doi.
org/10.1038/s41559-018-0777-y
Taylor, S. A., White, T. A., Hochachka, W. M., Ferretti, V., Curry, R.
L., & Lovette, I. (2014). Climate-mediated movement of an avian
hybrid zone. Current Biology: CB, 24(6), 671–676. https://doi.
org/10.1016/j.cub.2014.01.069
Trigo, T. C., Schneider, A., de Oliveira, T. G., Lehugeur, L. M., Silveira,
L., Freitas, T. R. O., & Eizirik, E. (2013). Molecular data reveal
complex hybridization and a cryptic species of neotropical wild cat.
Current Biology, 23(24), 2528–2533. https://doi.org/10.1016/j.
cub.2013.10.046
Turelli, M., & Orr, H. A. (2000). Dominance, epistasis and the genetics
of postzygotic isolation. Genetics, 154(4), 1663–1679. https://doi.
org/10.1093/genetics/154.4.1663. http://www.ncbi.nlm.nih.gov/
pubmed/10747061
Walsh, J., Billerman, S. M., Rohwer, V. G., Butcher, B. G., & Lovette, I.
J. (2020). Genomic and plumage variation across the controversial
Baltimore and Bullock’s oriole hybrid zone. The Auk, 137(4), 1–15.
https://doi.org/10.1093/AUK/UKAA044
Walsh, J., Kovach, A. I., Olsen, B. J., Shriver, W. G., & Lovette, I. J.
(2018). Bidirectional adaptive introgression between two ecolog-
ically divergent sparrow species. Evolution, 72(10), 2076–2089.
https://doi.org/10.1111/evo.13581
Wang, N., Borrell, J. S., Bodles, W. J. A., Kuttapitiya, A., Nichols, R.
A., & Buggs, R. J. A. (2014). Molecular footprints of the Holo-
cene retreat of dwarf birch in Britain. Molecular Ecology, 23(11),
2771–2782. https://doi.org/10.1111/mec.12768
Wang, S., Rohwer, S., Delmore, K., & Irwin, D. E. (2019). Crossdecades
stability of an avian hybrid zone. Journal of Evolutionary Biology,
32(11), 1242–1251. https://doi.org/10.1111/jeb.13524
Wielstra, B. (2019). Historical hybrid zone movement: More pervasive
than appreciated. Journal of Biogeography, 46(7), 1300–1305.
https://doi.org/10.1111/jbi.13600
Wielstra, B., Burke, T., Butlin, R. K., & Arntzen, J. W. (2017). A signature
of dynamic biogeography: Enclaves indicate past species replace-
ment. Proceedings of the Royal Society B: Biological Sciences,
284(1868), 20172014. https://doi.org/10.1098/rspb.2017.2014
Wringe, B. F., Stanley, R. R. E., Jeffery, N. W., Anderson, E. C., &
Bradbury, I. R. (2017). hybriddetective: A workow and package
to facilitate the detection of hybridization using genomic data in
R. Molecular Ecology Resources, 17(6), e275–e284. https://doi.
org/10.1111/1755-0998.12704
Yang, W., Feiner, N., Salvi, D., Laakkonen, H., Jablonski, D., Pinho, C.,
Uller, T., Sacchi, R., Zuf, M. A. L., Scali, S., Plavos, K., Palis, P.,
Poulakakis, N., Lymberakis, P., Jandzik, D., Schulte, U., Aubret, F.,
Badiane, A., Perezi de Lanuza, G., … Uller, T. (2022). Population
genomics of wall lizards reects the dynamic history of the Med-
iterranean basin. Molecular Biology and Evolution, 39(1), 2021.
https://doi.org/10.1093/MOLBEV/MSAB311
Yang, W., While, G. M., Laakkonen, H., Sacchi, R., Zuf, M. A. L., Scali,
S., Uller, T., & Uller, T. (2018). Genomic evidence for asymmetric
introgression by sexual selection in the common wall lizard. Molecu-
lar Ecology, 27(21), 4213–4224. https://doi.org/10.1111/mec.14861
Yuri, T., Jernigan, R. W., Brumeld, R. T., Bhagabati, N. K., & Braun,
M. J. (2009). The effect of marker choice on estimated levels of
introgression across an avian (Pipridae: Manacus) hybrid zone.
Molecular Ecology, 18(23), 4888–4903. https://doi.org/10.1111/
j.1365-294X.2009.04381.x
Zohren, J., Wang, N., Kardailsky, I., Borrell, J. S., Joecker, A., Nichols,
R. A., & Buggs, R. J. A. (2016). Unidirectional diploid–tetraploid
introgression among British birch trees with shifting ranges shown
by restriction siteassociated markers. Molecular Ecology, 25(11),
2413–2426. https://doi.org/10.1111/mec.13644
Downloaded from https://academic.oup.com/evolut/advance-article/doi/10.1093/evolut/qpae076/7675325 by guest on 13 June 2024
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Rivers frequently delimit the geographic ranges of species in the Amazon Basin. These rivers also define the boundaries between genetic clusters within many species, yet river boundaries have been documented to break down in headwater regions where rivers are narrower. To explore the evolutionary implications of headwater contact zones in Amazonia, we examined genetic variation in the Blue-capped Manakin (Lepidothrix coronata), a species previously shown to contain several genetically and phenotypically distinct populations across the western Amazon Basin. We collected restriction site-associated DNA sequence data (RADcap) for 706 individuals and found that spatial patterns of genetic structure indicate several rivers, particularly the Amazon and Ucayali, are dispersal barriers for L. coronata. We also found evidence that genetic connectivity is elevated across several headwater regions, highlighting the importance of headwater gene flow for models of Amazonian diversification. The headwater region of the Ucayali River provided a notable exception to findings of headwater gene flow by harboring non-admixed populations of L. coronata on opposite sides of a < 1-km-wide river channel with a known dynamic history, suggesting that additional prezygotic barriers may be limiting gene flow in this region.
Article
Full-text available
Sexual selection by mate choice is a powerful force that can lead to evolutionary change, and models of why females choose particular mates are central to understanding its effects. Predominant mate choice theories assume preferences are determined solely by genetic inheritance, an assumption still lacking widespread support. Moreover, preferences often vary among individuals or populations, fail to correspond with conspicuous male traits, or change with context, patterns not predicted by dominant models. Here, we propose a new model that explains this mate choice complexity with one general hypothesized mechanism, “Inferred Attractiveness.” In this model, females acquire mating preferences by observing others’ choices and use context-dependent information to infer which traits are attractive. They learn to prefer the feature of a chosen male that most distinguishes him from other available males. Over generations, this process produces repeated population-level switches in preference and maintains male trait variation. When viability selection is strong, Inferred Attractiveness produces population-wide adaptive preferences superficially resembling “good genes.” However, it results in widespread preference variation or nonadaptive preferences under other predictable circumstances. By casting the female brain as the central selective agent, Inferred Attractiveness captures novel and dynamic aspects of sexual selection and reconciles inconsistencies between mate choice theory and observed behavior.
Article
Full-text available
Amazonian rivers represent known barriers for avian dispersal, reducing gene flow and enhancing differentiation. Despite the importance of rivers in the avian evolutionary process, we have made only minor advances in understanding the limitations imposed by rivers on flying birds. To fill that gap, we conducted dispersal-challenge experiments over water, assessing the flying capabilities of 84 tropical bird species of 22 different avian families. We mist-netted and released 484 birds from a stationary boat on the Rio Branco, northern Amazonia, at increasing distances from the shore, including 249 individuals at 100; 219 at 200; 8 at 300; and 5 at 400 m. A successful trial was represented by a bird reaching the riverbank, whereas a failure would refer to birds not reaching the shore and landing on the water, when they were rescued by our team. Our main goal was to understand if the outcome in the experiments could be predicted by (i) phylogenetic constraints, (ii) morphology (body mass and wing shape), (iii) flight speed, (iv) ecological preferences (stratum, habitat, and river-island specialization), and (v) psychological reluctance to fly. Nearly two thirds of the individuals (332) were successful in reaching the riverbank, whereas 152 failed. We found significant differences among lineages. Whereas seven avian families succeeded in all of the trials, two families (antbirds and wrens) were particularly bad dispersers (<40% success). The hand-wing index (HWI) was the single most powerful predictor of trial success. Flying speed was also a significant predictor of success. Overall, ecological attributes had a low explanatory power. Only forest stratum preference had a significant, although weak, effect on dispersal ability: canopy- and ground-dwellers performed better than understory birds. However, we found no effect of habitat preference or river-island specialization on dispersal ability. Our speed estimates for 64 bird species are among the first produced for the tropics and suggest slower flying speeds than those reported from temperate migratory birds. Although birds showed behavioral differences when presented with the opportunity to fly away from the boat, we found no evidence that their reluctance to fly could predict the outcome in the experiments. This represents the first experimental study evaluating the riverine effect through dispersal ability of Amazonian birds, providing important insights to better understand dispersal limitations provided by riverine barriers.
Article
Full-text available
The Mediterranean Basin has experienced extensive change in geology and climate over the past six million years. Yet, the relative importance of key geological events for the distribution and genetic structure of the Mediterranean fauna remains poorly understood. Here, we use population genomic and phylogenomic analyses to establish the evolutionary history and genetic structure of common wall lizards (Podarcis muralis). This species is particularly informative because, in contrast to other Mediterranean lizards, it is widespread across the Iberian, Italian, and Balkan peninsulas, and in extra-Mediterranean regions. We found strong support for six major lineages within P. muralis, which were largely discordant with the phylogenetic relationship of mitochondrial DNA. The most recent common ancestor of extant P. muralis was likely distributed in the Italian Peninsula, and experienced an “Out-of-Italy” expansion following the Messinian salinity crisis (∼5 Mya), resulting in the differentiation into the extant lineages on the Iberian, Italian and Balkan peninsulas. Introgression analysis revealed that both inter- and intraspecific gene flows have been pervasive throughout the evolutionary history of P. muralis. For example, the Southern Italy lineage has a hybrid origin, formed through admixture between the Central Italy lineage and an ancient lineage that was the sister to all other P. muralis. More recent genetic differentiation is associated with the onset of the Quaternary glaciations, which influenced population dynamics and genetic diversity of contemporary lineages. These results demonstrate the pervasive role of Mediterranean geology and climate for the evolutionary history and population genetic structure of extant species.
Article
Full-text available
Hybrid zones provide a window into the evolutionary processes governing species divergence. Yet, the contribution of mate choice to the temporal and spatial stability of hybrid zones remains poorly explored. Here, we investigate the effects of assortative mating on hybrid zone dynamics by means of a mathematical model parameterized with phenotype and genotype data from the hybrid zone between all-black carrion and grey-coated hooded crows. In the best-fit model, narrow clines of the two mating-trait loci were maintained by a moderate degree of assortative mating inducing pre- and post-zygotic isolation via positive frequency-dependent selection. Epistasis between the two loci induced hybrid-zone movement in favor of alleles conveying dark plumage followed by a shift in the opposite direction favouring grey-coated phenotypes 1,200 generations after secondary contact. Unlinked neutral loci diffused near-unimpeded across the zone. These results were generally robust to the choice of matching rule (self-referencing or parental imprinting) and effects of genetic drift. Overall, this study illustrates under which conditions assortative mating can maintain steep clines in mating-trait loci without generalizing to genome-wide reproductive isolation. It further emphasizes the importance of mating-trait architecture for spatio–temporal hybrid-zone dynamics. This article is protected by copyright. All rights reserved
Article
Full-text available
In the past decade, advances in genome sequencing have allowed researchers to uncover the history of hybridization in diverse groups of species, including our own. Although the field has made impressive progress in documenting the extent of natural hybridization, both historical and recent, there are still many unanswered questions about its genetic and evolutionary consequences. Recent work has suggested that the outcomes of hybridization in the genome may be in part predictable, but many open questions about the nature of selection on hybrids and the biological variables that shape such selection have hampered progress in this area. We synthesize what is known about the mechanisms that drive changes in ancestry in the genome after hybridization, highlight major unresolved questions, and discuss their implications for the predictability of genome evolution after hybridization.
Article
Full-text available
Brightly colored manakin (Aves: Pipridae) males are known for performing acrobatic displays punctuated by non-vocal sounds (sonations) in order to attract dull colored females. The complexity of the display sequence and assortment of display elements involved (e.g., sonations, acrobatic maneuvers, and cooperative performances) varies considerably across manakin species. Species-specific display elements coevolve with display-distinct specializations of the neuroanatomical, muscular, endocrine, cardiovascular, and skeletal systems in the handful of species studied. Conducting a broader comparative study, we previously found positive associations between display complexity and both brain mass and body mass across 8 manakin genera, indicating selection for neural and somatic expansion to accommodate display elaboration. Whether this gross morphological variation is due to overall brain and body mass expansion (concerted evolution) versus size increases in only functionally relevant brain regions and growth of particular body (“somatic”) features (mosaic evolution) remains to be explored. Here we test the hypothesis that cross-species variation in male brain mass and body mass is driven by mosaic evolution. We predicted positive associations between display complexity and variation in the volume of the cerebellum and sensorimotor arcopallium, brain regions which have roles in sensorimotor processes, and learning and performance of precisely timed and sequenced thoughts and movements, respectively. In contrast, we predicted no associations between the volume of a limbic arcopallial nucleus or a visual thalamic nucleus and display complexity as these regions have no-specific functional relationship to display behavior. For somatic features, we predicted that the relationship between body mass and complexity would not include contributions of tarsus length based on a recent study suggesting selection on tarsus length is less labile than body mass. We tested our hypotheses in males from 12 manakin species and a closely related flycatcher. Our analyses support mosaic evolution of neural and somatic features functionally relevant to display and indicate sexual selection for acrobatic complexity may increase the capacity for procedural learning via cerebellar enlargement and maneuverability via a reduction in tarsus length in species with lower overall complexity scores.
Article
A fundamental goal in evolutionary biology is to understand the genetic architecture of adaptive traits. Using whole-genome data of 3955 of Darwin’s finches on the Galápagos Island of Daphne Major, we identified six loci of large effect that explain 45% of the variation in the highly heritable beak size of Geospiza fortis, a key ecological trait. The major locus is a supergene comprising four genes. Abrupt changes in allele frequencies at the loci accompanied a strong change in beak size caused by natural selection during a drought. A gradual change in Geospiza scandens occurred across 30 years as a result of introgressive hybridization with G. fortis . This study shows how a few loci with large effect on a fitness-related trait contribute to the genetic potential for rapid adaptive radiation.
Article
The ability to move across the landscape is a fundamental property of species that can determine their chances of persistence in fragmented landscapes and in rapidly changing environments. Despite its importance, empirical evidence showing the effect of movement capacity on patterns of movement across fragmented landscapes is limited. In this study, we examine the role of flight efficiency on the likelihood of crossing a man‐made habitat gap. We used data from the Biological Dynamics of Forest Fragments Projects on recaptures of banded birds in an Amazonian forest bisected by a road. For a total of 45 species, we estimated flight efficiency using the hand‐wing index (a proxy for the wing's aspect ratio) and used it as a predictor of the probability of road crossing in phylogenetic binomial regression models. We found that flight efficiency was a strong predictor of road‐crossing probability: species with high hand‐wing indices crossed the road more frequently than those with low hand‐wing indices. In contrast, other characteristics such as body mass, diet, flocking behavior, and foraging stratum did not show significant associations with road‐crossing probability. Our results suggest that proxies of flight efficiency such as the hand‐wing index can be powerful tools for predicting the vulnerability of bird species to forest fragmentation. Abstract in Spanish is available with online material
Article
Natural hybrid zones have provided important insights into the evolutionary process, and their geographic dynamics over time can help to disentangle the underlying biological processes that maintain them. Here, we leverage replicated sampling of an identical transect across the hybrid zone between yellow-shafted and red-shafted flickers in the Great Plains to assess its stability over ∼60 years (1955-1957 to 2016-2018). We identify a ∼73 km westward shift in the hybrid zone center towards the range of the red-shafted flicker, but find no associated changes in width over our sampling period. In fact, the hybrid zone remains remarkably narrow, suggesting some kind of selective pressure maintains the zone. By comparing to previous work in the same geographic region, it appears likely that the movement in the hybrid zone has occurred in the years since the early 1980s. This recent movement may be related to changes in climate or land management practices that have allowed westward movement of yellow-shafted flickers into the Great Plains. This article is protected by copyright. All rights reserved.