Sie sind auf Seite 1von 10

Systematics and Biodiversity

ISSN: 1477-2000 (Print) 1478-0933 (Online) Journal homepage: http://www.tandfonline.com/loi/tsab20

Relationships of morphological groups in the


northern flicker superspecies complex (Colaptes
auratus & C. chrysoides)

Joseph D. Manthey, Mark Geiger & Robert G. Moyle

To cite this article: Joseph D. Manthey, Mark Geiger & Robert G. Moyle (2016): Relationships
of morphological groups in the northern flicker superspecies complex (Colaptes auratus & C.
chrysoides), Systematics and Biodiversity, DOI: 10.1080/14772000.2016.1238020

To link to this article: http://dx.doi.org/10.1080/14772000.2016.1238020

View supplementary material

Published online: 20 Oct 2016.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tsab20

Download by: [Ryerson University Library] Date: 21 October 2016, At: 10:44
Systematics and Biodiversity (2016), 19

Research Article
Relationships of morphological groups in the northern flicker
superspecies complex (Colaptes auratus & C. chrysoides)

1,3
JOSEPH D. MANTHEY , MARK GEIGER2 & ROBERT G. MOYLE 1

1
Biodiversity Institute and Department of Ecology and Evolutionary Biology, University of Kansas, Lawrence, KS 66045, USA
2
Palumbo-Donahue School of Business, Duquesne University, Pittsburgh, PA 15282, USA
3
Biology Department, New York University Abu Dhabi, Abu Dhabi, UAE
(Received 30 May 2016; accepted 7 September 2016)

Many species complexes have diversified rapidly and recently, resulting in morphologically diverse populations; however,
the rapid pace of diversification often prevents identification of clear phylogeographic structure. Recently, the use of large
genomic and reduced-representation genomic datasets has improved resolution of the evolutionary histories in such species
and allowed identification of lineages on distinct evolutionary trajectories. The northern flicker (Colaptes auratus) and
gilded flicker (Colaptes chrysoides) form a polytypic superspecies group with a complex taxonomic history. The
superspecies group includes up to 13 described subspecies, which represent slight geographic variation among five main
morphological groups: red-shafted flickers of western North America (cafer group), yellow-shafted flickers of eastern
North America (auratus group), Cuban flickers of the Caribbean (chrysocaulosus group), gilded flickers of the U.S. south-
west and Mexican north-west (chrysoides group), and Guatemalan flickers of Central America (mexicanoides group).
These groups are largely differentiable by variation in feather shaft colour, malar colour, throat colour, crown colour, and
back barring. Here, using mitochondrial DNA (mtDNA) and hundreds of single nucleotide polymorphisms (SNPs), we
characterized the genetic relationships and genomic distinctiveness of the five morphological groups. We found the
mexicanoides group to be the most genetically distinct in both mtDNA (1.4% sequence divergence) and large SNP
panels. The chrysocaulosus group is differentiated by a single basepair mutation in a small mtDNA fragment. In both
mtDNA and SNP panels, there is little genetic distinctiveness between auratus, cafer, and chrysoides morphological
groups, with evidence of admixture and a lack of fixed differences.
Key words: admixture, birds, Colaptes, Picidae, phylogeography, RAD-seq, species tree, woodpeckers

Introduction evolved distinct phenotypic forms. For example, the


Dark-eyed Junco (Junco hyemalis) has six distinct mor-
Avian species range from monomorphic to highly poly-
phological forms, but they lack identifiable genetic differ-
typic. Exemplar polytypic taxa include some North Ameri-
entiation (McCormack et al., 2012; Mila, McCormack,
can sparrows (e.g. Song Sparrow (Melospiza melodia),
Casta~neda, Wayne, & Smith, 2007). The nature of recent
Zink, 2010), with many phenotypically distinct subspecies.
divergence events necessitates the use of larger genetic
Perhaps the most notable example, the Australasian Golden
datasets for phylogeographic and phylogenetic inference.
Whistler species complex (Pachycephala melanura/pector-
Using large genetic datasets allows improved resolution
alis; Andersen et al., 2014), includes more than 70
for identifying species rapid and recent evolutionary his-
described subspecies, many of which have diagnostic phe-
tories, as well as identifying lineages on distinct evolu-
notypic characters and isolated geographic ranges.
tionary trajectories (e.g., evolutionary significant units).
A shift toward the evolutionary species concept (ESC;
Restriction-site associated DNA sequencing (RAD-seq;
Wiley & Mayden, 2000) has led to a greater reliance on
Miller, Dunham, Amores, Cresko, & Johnson, 2007) has
genetics, ecology, and geography for defining species lim-
emerged as a useful method for obtaining large panels of
its (Chesser et al., 2015). However, since the glacial peri-
single nucleotide polymorphisms (SNPs), which has been
ods of the Pleistocene, many avian species have rapidly
valuable for identifying recent genetic differentiation in a
number of organisms, including multiple species of birds
Correspondence to: Joseph D. Manthey. E-mail: (McCormack et al., 2012).
jdmanthey@gmail.com

ISSN 1477-2000 print / 1478-0933 online


The Trustees of the Natural History Museum, London 2016. All Rights Reserved.
http://dx.doi.org/10.1080/14772000.2016.1238020
2 J. D. Manthey et al.

The northern flicker (Colaptes auratus) and gilded barriers, respectively. Although geographic phenotypic
flicker (Colaptes chrysoides) form a polytypic superspe- variation and hybridization between yellow- and red-
cies group (Winkler & Christie, 2002) with a complex his- shafted flickers have been characterized (Grudzien,
tory of taxonomic splitting and lumping (Monroe et al., Moore, Cook, & Tagle, 1987; Wiebe & Bortolotti, 2001,
1995; Moore, Graham, & Price, 1991), limited evidence 2002), the genetic structure of the superspecies complex
of genetic differentiation (Fletcher & Moore, 1992; Moore has not been investigated. While widespread mitochon-
et al., 1991), and varying numbers of currently recognized drial DNA (mtDNA) haplotype sharing is evident among
species by different resources (e.g., Chesser et al., 2015; yellow- and red-shafted flickers (Moore et al., 1991), little
del Hoyo et al., 2016). The superspecies group contains allozyme genetic structure is apparent among the two
up to 13 described subspecies, which represent geographic (Fletcher & Moore, 1992), and weak genetic differentia-
variation of five main morphological groups (Fig. 1; tion exists between gilded flicker and northern flicker
Short, 1982): red-shafted flickers of western North Amer- (cafer and auratus groups) individuals (Moore et al.,
ica (cafer group), yellow-shafted flickers of eastern North 1991), no studies have included genetic data of two of the
America (auratus group), Cuban flickers of the Caribbean five morphological groups (Cuban and Guatemalan).
(chrysocaulosus group), gilded flickers of the U.S. south- Additionally, as part of a large supermatrix investigation
west and Mexican north-west (chrysoides group), and of Picidae, Dufort (2016) found C. auratus sister to C.
Guatemalan flickers of Central America (mexicanoides chrysoides, and not C. cafer, the taxon with which it
group). These groups are largely differentiable by varia- hybridizes, albeit with little support and warranting fur-
tion in feather shaft colour, malar colour, throat colour, ther investigation.
crown colour, and back barring. To characterize the phylogenetic placement of the five
Where morphological groups come into contact, there morphological groups in the Colaptes auratus superspe-
is extensive interbreeding (Short, 1982) with hybridiza- cies complex, we created two datasets: (1) we sequenced
tion between yellow- and red-shafted flickers most wide- a mtDNA marker from each of the morphological groups;
spread (Fig. 1; McCarthy, 2006). However, two and (2) we used a restriction site associated DNA
morphological groups, the Cuban and Guatemalan flick- sequencing (RAD-seq) method to obtain hundreds of sin-
ers, are allopatric from all other taxa, with the Florida gle nucleotide polymorphisms (SNPs) from four of the
Strait and Isthmus of Tehuantepec as biogeographic five morphological groups (excluding the Cuban group
due to lack of modern genetic samples). With these data-
sets, our goal was to infer the phylogenetic relationships
of the five morphological groups and assess genetic dis-
tinctiveness of each group.

Materials and methods


Sampling, laboratory procedures, and SNP
dataset creation
Fresh tissue samples of 13 Colaptes auratus and Colaptes
chrysoides individuals, representing four of the five major
morphological groups of the Colaptes auratus species
complex, were obtained from across North and Central
America (Table 1) and from localities outside of putative
hybrid zones. We obtained two toe-pad cuttings of
Colaptes auratus chrysocaulosus, representing the fifth
morphological group of the Colaptes auratus complex, as
no fresh tissue samples are currently available from Cuba.
Two tissue samples of Colaptes rubiginosus were
obtained from one locality in South America (Table 1) for
use as outgroup samples. For all fresh tissue samples, total
genomic DNA was extracted using a QIAGEN DNeasy
tissue extraction kit following manufacturer protocols. To
Fig. 1. Approximate breeding-range distributions for each of the extract DNA from toe-pad samples, we used the same
five morphological groups in the Northern Flicker superspecies methodology as fresh tissue following an initial overnight
complex, including the hybrid-zone region between the cafer
and auratus groups. The chrysocaulosus and mexicanoides digestion including dithiothreitol (1M).
groups are emphasized in boxes for clarity. The map is projected For all fresh tissue samples, we obtained the entire
using the Albers Equal Area Conical North America Projection. mitochondrial ND2 gene using previously published
Northern flicker RAD-seq relationships 3

primers L5215 (Hackett, 1996) and H6313 (Johnson & the original settings for all subsequent analyses. This pipe-
Sorenson, 1998). For toe-pad samples, we were able to line resulted in two SNP datasets, differing in the minimum
sequence 231 bp of the ND2 gene using L5215 and a cus- sequence quality threshold (Phred score of Q10 or Q30).
tom PCR primer (CTgTggATCAggCATTgATTA), as These two datasets were used in all subsequent analyses.
other custom primers were unable to successfully amplify
additional regions of ND2. All 10 mL PCR reactions were
subjected to an initial denaturation period of 10 min at
Investigation of genetic structure and
95 C, with 35 subsequent cycles of 95 C for 30 s, 54 C phylogenetic relationships
for 45 s, and 72 C for 1 min. All PCR products were puri- With the ND2 data, we created median-joining networks
fied using Exosap-IT (USB Corporation), cycle-sequenced (Bandelt, Forster, & Rohl, 1999) to visualize relationships
using 7 mL ABI BigDye (Applied Biosystems) reactions, among haplotypes using PopART (popart.otago.ac.nz).
purified using a standard ethanol cleanup, and sequenced Additionally, we created phylogenetic trees using PhyML
(both strands) on an ABI Genetic Analyzer. All sequences (Guindon & Gascuel 2003) with an HKY model of
were manually aligned and quality-checked using sequence evolution (based on model selection chosen
Sequencher v4.8 (GeneCodes). with a Bayesian information criterion) and 1000 bootstrap
We performed a modified RAD-seq (Miller et al., 2007) replicates to assess support.
protocol identical to the methods used by Manthey and To investigate genetic structure among individuals
Moyle (2015) to obtain SNP data for all individuals with without a priori population assignment, we used the pro-
fresh tissue samples. Briefly, we used the restriction gram STRUCTURE (Pritchard, Stephens, & Donnelly,
enzyme NdeI to digest samples, multiplexed with one bar- 2000) with the SNP dataset. Initially, we inferred lambda
code per individual, and used a Pippin Prep electrophore- with the number of inferred populations (k) restricted to
sis cassette (Sage Science) to size select fragments k D 1. Subsequently, we used a fixed lambda, as identified
between 500 and 600 bp. Following DNA quality and in the first run, with a burn-in period of 50 000 steps fol-
quantity tests using quantitative PCR and an Agilent lowed by 150 000 MCMC iterations; using these settings,
Tapestation, all samples were sequenced on a partial lane we ran five replicates of STRUCTURE with k D 15. To
of an Illumina HiSeq2500 (100 bp single-end reads) at the identify the number of inferred populations, we used the
University of Kansas Genome Sequencing Core Facility. DK method of Evanno, Regnaut, and Goudet (2005).
To assemble loci de novo from the Illumina sequencing To further investigate genetic structure among morpho-
run data files, we used the STACKS (Catchen, Amores, logical groups, we used discriminant analysis of principal
Hohenlohe, Cresko, & Postlethwait, 2011) pipeline. The components (DAPC; Jombart, Devillard, & Balloux, 2010)
included python script process_RADtags was used to with the R package adegenet (Jombart & Ahmed, 2011).
quality-screen sequences, including eliminating those DAPC uses principal components analysis followed by dis-
with possible adapter contamination or lacking the restric- criminant analysis to identify genetic clusters within the
tion site. We used a quality threshold in process_RADtags data. We used the Bayesian information criterion to obtain
of an average Phred score of 10 in sliding windows of estimates of the optimal number of genetic clusters.
15 bp. To assess the effects of this threshold on down- Because the STRUCTURE analyses identified some
stream analyses, we repeated this step with a higher qual- individuals that appeared to be admixed (see Results), we
ity threshold (Phred score of 30), and performed library inferred relationships among morphological groups using
construction and downstream analyses. The ustacks, TreeMix (Pickrell & Pritchard, 2012). TreeMix infers a
cstacks, and sstacks modules of STACKS were used to maximum likelihood phylogeny using covariance of allele
create SNP libraries for each individual. Lastly, the popu- frequencies between populations, followed by identifying
lations module of STACKS was used to create final SNP groups that are more closely related than can be explained
datasets used for analysis. We used the following restric- by the phylogeny; these groups represent potential gene
tions to include SNPs: a minimum allele frequency of flow events (i.e., migration edges on the phylogeny;
0.05, a minimum stack depth of five, observed heterozy- Pickrell & Pritchard, 2012). We ran TreeMix with each of
gosity less than 0.5 to reduce inclusion of paralogous loci, the two SNP datasets, followed by 100 bootstrap repli-
and present in 50% of individuals. From this subset, we cates to assess confidence in phylogeny estimation. Boot-
selected loci present in a minimum of 75% of individuals straps used sets of 200 SNPs. Migration edges were added
from every morphological group (75% coverage matrix to the phylogeny until they explained 99.8% of the vari-
(CM)). To ensure that minimum stack depth did not ance in the SNP data (threshold used in other studies;
impact results, we reran the last step of STACKS with dif- Decker et al., 2014; Pickrell & Pritchard, 2012); this
ferent minimum stack depth values (m D 1, 5, 10, 20). resulted in no migration edges for either dataset.
Using these different settings did not change values of In addition to TreeMix, we estimated a species tree using
genetic differentiation (FST) among morphological groups SVDquartets (Chifman & Kubatko, 2014), which infers the
(r > 0.979, p < 0.001); we therefore used the dataset with topologies of individual quartets in a coalescent framework,
4 J. D. Manthey et al.

followed by the use of those quartets to infer a species tree. variable (Table 1) with a mean of 702 066 reads (sd D 561
We exhaustively sampled quartets from the dataset and 469 reads). The final SNP datasets, the 75% CMs with dif-
inferred a species tree from all quartets. We implemented ferent minimum quality scores (hereafter Q10 and Q30),
SVDquartets using PAUP (Swofford, 2003). Because gene contained 1255 and 734 SNPs, respectively. Of the SNPs
flow among lineages would likely affect species tree esti- in the final Q10 dataset, the median sequencing coverage
mation (Leache, Harris, Rannala, & Yang, 2013), we was 24 reads (sd D 26.7). The majority of SNPs was either
removed admixed individuals (Table 1) prior to running private to a morphological group, or shared between
SVDquartets. groups (Fig. 3). Only the mexicanoides group had more
than a single fixed difference (Fig. 3). Most FST pairwise
Results comparisons among groups, with the exception of the
mexicanoides group, were low (Table 2). When admixed
mtDNA sequencing individuals were removed (see STRUCTURE results and
We recovered 231 bp of the ND2 gene for toepad samples, Table 1), FST pairwise comparisons were again highest
and the entire gene (1041 bp) for all other individuals. for comparisons with the mexicanoides group (Table 2).
Using only 231 bp, the auratus, cafer, and chrysoides indi- Analyses of genetic structure inferred various numbers
viduals all shared the same haplotype, both chrysocaulosus of genetic clusters, varying among datasets and analyses.
individuals shared the same haplotype, which was a single In DAPC, a single genetic cluster had the lowest BIC
bp different from the auratus/cafer/chrysoides haplotype, score, but all values between one and four were qualita-
and both mexicanoides individuals had the same haplotype tively similar. For STRUCTURE analyses, the DK
with three mutational differences from the most common method indicated two and three genetic clusters for differ-
haplotype (Fig. 2). Using the entire gene (and excluding ent datasets. Because of the varied outcomes, we present
chrysocaulosus), the mexicanoides individuals shared the results for all analyses with two, three, and four genetic
same haplotype and were 15 mutational steps (1.4% clusters (Fig. 4). Depending on the number of genetic
divergence) from the nearest other individual in the haplo- clusters assigned, two or three individuals (individuals 1,
type network (Fig. 1). In contrast, among the auratus, 7, 11 in Table 1 & Fig. 4) showed clear signs of admixture
cafer, and chrysoides samples, no individuals were more among morphological groups. The auratus and cafer indi-
than four mutational steps from each other (Fig. 2). Inter- viduals were generally indistinguishable unless the num-
estingly, some chrysoides haplotypes were more closely ber of genetic clusters was four in the Q30 dataset, though
related to cafer or auratus haplotypes than to each other.
Phylogenetic analyses using both the short and full align-
ments were rather uninformative, because the only well-
supported relationship was a sister relationship between the
two individuals of mexicanoides; all other nodes had low
00
1000
or no support, resulting in a comb-shaped phylogeny (See
online supplemental material, which is available from the
Frequency

articles Taylor & Francis Online page at http://dx.doi.org/ 00


100
10.1080/14772000.2016.1238020).

RAD-seq analyses 0
10
Following quality control, we obtained 10 530 990
sequencing reads from the 13 ingroup and two outgroup
samples. The number of reads per individual was highly
0
us

an s

s
fe

de
at

id
ca

oi
o
r
au

s
ry

auratus
ic
ch

231 bp of ND2
ex

cafer
chrysocaulosus
m

chrysoides 1041 bp of ND2


mexicanoides (w/o chrysocaulosus) = fixed = shared
= private
Fig. 2. Median-joining haplotype networks of the mitochondrial
DNA gene ND2. Circles are proportional to haplotype sample Fig. 3. Polymorphisms summary. Frequencies of fixed, shared,
size. Haplotypes are connected with lines and bars indicating the and private polymorphisms in the 75% coverage matrix (Q10)
number of mutational steps. dataset. Note the log scale.
Northern flicker RAD-seq relationships 5

Table 1. Samples used in this study (Voucher), morphological group or outgroup (Group), individual number in Figure 3 (Ind.), country
and state (Locality), number of quality-controlled RAD sequence reads used in study (# Reads), number of RAD-tag loci recovered for
each individual, coverage of individuals (Coverage; median and sd), identification of three highly admixed individuals (Adm.). Asterisks
() indicate latitude and longitude are approximated. Number of reads, RAD-tags, and coverage are all based on the 75% coverage
matrix (Q10) dataset, with the stats for the Q30 dataset in the supplementary information.

Voucher Group Lat. Long. Ind. Locality # Reads RAD-tags Coverage Adm

KU 25268 auratus 39.035 94.574 1 USA: Missouri 772,265 31,574 9 (4.0) X


KU 30947 auratus 39.036 94.578 2 USA: Missouri 1,071,004 43,399 41 (14.2)
KU 30060 cafer 33.811 109.159 3 USA: Arizona 666,655 31,386 38 (14.3)
KU 30071 cafer 32.419 110.735 4 USA: Arizona 936,826 37,438 41 (15.5)
KU 30083 cafer 34.447 111.359 5 USA: Arizona 1,181,226 34,111 44 (16.2)
KU 30095 cafer 36.461 112.120 6 USA: Arizona 395,670 15,267 31 (13.2)
KU 30098 cafer 36.461 112.120 7 USA: Arizona 835,217 30,699 31 (10.6) X
LSUMZ B-63594 chrysocaulosus 21.7 77.9 Cuba: Camaguay
LSUMZ B-63595 chrysocaulosus 22.3 81.2 Cuba: Villa Clara
KU 30078 chrysoides 32.344 112.893 8 USA: Arizona 2,278,487 42,234 102 (42.0)
KU 30079 chrysoides 32.344 112.893 9 USA: Arizona 83,673 7,172 6 (2.7)
UWBM 84071 chrysoides 26.310 108.810 10 Mexico: Sinaloa 255,796 14,002 19 (8.0)
UWBM 90812 chrysoides 26.305 108.808 11 Mexico: Sinaloa 839,849 29,809 15 (7.1) X
KU 4974 mexicanoides 13.991 88.085 12 El Salvador: Morazon 160,322 11,266 12 (5.1)
KU 4986 mexicanoides 13.991 88.085 13 El Salvador: Morazon 322,974 18,451 20 (7.2)
KU 3926 C. rubiginosus 5.283 60.750 Guyana: Cuyuni-Mazaruni 615,976 28,089 37 (11.6)
KU 4043 C. rubiginosus 5.267 60.733 Guyana: Cuyuni-Mazaruni 115,050 9,963 8 (3.1)

this signal was not in the Q10 dataset (Fig. 4). The stron- genetic distinctiveness of the mexicanoides group using
gest signal of genetic structure was between the mexica- both mtDNA and SNPs. Relationships of the auratus,
noides group and all other groups, with some evidence of cafer, chrysoides, and chrysocaulosus groups are mixed
distinctiveness of the chrysoides group. or unresolved in various datasets and analyses. Although
In the phylogenetic analyses TreeMix and SVDquartets, this study does not include comprehensive sampling
the mexicanoides group was identified as sister to all other across the entire distribution of this complex, it provides
groups, with the remaining relationships conflicting the first step toward understanding the evolutionary his-
between analyses (Fig. 5). In TreeMix, mixed relationships tory of the entire superspecies group; the results suggest
among chrysoides, cafer, and auratus varied in analysis of that hundreds of SNPs may not be enough to resolve the
different datasets (not shown). The SVDquartets analyses recent evolutionary history of the species complex, war-
identified a sister relationship of chrysoides and auratus C ranting a much larger number of SNPs or whole genome
cafer, though this was not strongly supported (Fig. 5). sequencing across a greater geographic breadth of popu-
lations in future studies.

Discussion
Here, we present the first genetic data for two of the five The Isthmus of Tehuantepec
morphological groups in the northern flicker superspecies biogeographic barrier
complex (C. auratus and C. chrysoides). We found
This study identified two major biogeographic barriers
Table 2. Pairwise FST values for reduced and full datasets, above separating populations: the Florida Strait separates Cuban
and below the diagonal, respectively. Individuals labelled as and U.S. mainland populations, and the Isthmus of
admixed in Table 1 were removed for the reduced dataset. Values Tehuantepec separates Central American and more north-
are reported with the format: 75% coverage matrix Q10/Q30. ern populations. The Isthmus of Tehuantepec is a low ele-
auratus cafer chrysoides mexicanoides vation valley separating montane habitats in central
Mexico from those of southern Mexico and Central Amer-
auratus 0.154/0.154 0.268/0.286 0.605/0.581 ica. The northern flicker (auratus, cafer, chrysocaulosus,
cafer 0.099/0.098 0.184/0.192 0.294/0.281 and mexicanoides groups) is largely a generalist of open
chrysoides 0.170/0.168 0.131/0.130 0.393/0.405 woodlands across its range (del Hoyo et al., 2016). How-
mexicanoides 0.338/0.308 0.173/0.165 0.291/0.285 ever, in the cafer and mexicanoides groups, this habitat
generally tends to be montane and riverine forest in the
6 J. D. Manthey et al.

k=2 k=3 k=4 the last two million years, consistent with the recent diver-
13
mexicanoides gence of Colaptes auratus mexicanoides from more north-
Quality Min. Q10

12

ern populations (based on 1.4% mtDNA sequence


(1255 SNPs)

11
10
chrysoides
9
8 divergence). These two comparative studies did not agree
7
6
5 cafer
on modes of diversification. Barber and Klicka (2010)
4
3
suggested that Pleistocene climate cycles, and possibly
2
1
auratus late Pliocene sea level changes, influenced diversification
in the region. Alternatively, Ornelas and colleagues
13
mexicanoides
(2013) suggested species-specific vicariance events
Quality Min. Q30

12
11
(734 SNPs)

10
9
chrysoides because of their inference of non-simultaneous divergen-
8
7
ces among species. More geographic sampling of the red-
6
5 cafer shafted flickers (cafer group) and Guatemalan flickers
4
3 (mexicanoides group) is warranted to definitively deter-
2
1
auratus
mine mechanisms of diversification between these groups.

Fig. 4. Genetic structure of morphological groups using SNP


data. (A) STRUCTURE and DAPC results for the 75% coverage Genetic distinctiveness of the morphological
matrices (CMs; Q10 and Q30) with two to four genetic clusters groups
(k D 2, 3, 4). Each row indicates probability of assignment to
genetic clusters (STRUCTURE); next to each row is the genetic Previously published DNA sequence data of C. chrysoides
cluster assignment (small box) from DAPC analyses. Numbered (Dufort, 2016), including two nuclear genes (myoglobin
labels (far right) refer to individuals in Table 1. and transforming growth factor beta-2; as of August 2016
on NCBIs GenBank) from a single individual, only had
two distinct polymorphisms from northern flicker (auratus
western U.S., Mexico, and Central America. These quali- or cafer groups) samples (< 0.2% sequence divergence).
ties of the species groups habitat in the region allows the This is similar to the lack of variation found in a previous
Isthmus of Tehuantepec to act as a biogeographic barrier. study of mtDNA differences between chrysoides and red-
In birds, the Isthmus of Tehuantepec barrier has been shafted cafer (Moore et al., 1991). Here, we found the
demonstrated in many Passeriformes species (Arbelaez- chrysoides mtDNA haplotypes to be non-monophyletic
Cortes, Nyari, & Navarro-Sig uenza, 2010; Barber & (Fig. 2) and the SNP variation did not clearly distinguish
Klicka, 2010; Barrera-Guzman, Mila, Sanchez-Gonzalez, the chrysoides individuals (Fig. 4), although there was
& Navarro-Sig uenza, 2012; Manthey, Klicka, & Spell- limited signal of distinctiveness when the number of
man, 2011) as well as some woodpeckers (Klicka, Spell- genetic clusters was three or more, especially in the Q30
man, Winker, Chua, & Smith, 2011) and hummingbirds dataset (Fig. 4).
(Barber & Klicka, 2010; Ornelas et al., 2013). In two com- Similarly, limited evidence of genetic differentiation
parative analyses (Barber & Klicka, 2010; Ornelas et al., exists between red-shafted cafer and yellow-shafted aura-
2013), including birds, mammals, and plants, diversifica- tus in previous work (Dufort, 2016; Fletcher & Moore,
tion across the Isthmus of Tehuantepec occurred within 1992; Moore et al., 1991). This is probably due to high
levels of gene flow between the two morphological
groups, which is suggested based on both banding-recov-
auratus
55/61 ery data of the two groups and conditional allele fre-
quency curves across the auratus-cafer hybrid zone
cafer (Grudzien et al., 1987). Because the cafer form hybridizes
100/97 with both the chrysoides and auratus forms (more fre-
quently with the latter) there may be sufficient gene flow
chrysoides to maintain only low levels of differentiation throughout
their genomes. This raises the possibility that relatively
small genomic regions maintain the phenotypic, behav-
mexicanoides ioural, and ecological differences of these three morpho-
logical groups. Genomic islands of speciation have been
proposed in other hybridizing bird species (e.g., Ficedula
C. rubiginosus flycatchers, Ellegren et al., 2012) where specific genomic
regions show increased levels of genetic differentiation
Fig. 5. Results of SVDquartets species tree analyses. Support on
each branch is indicated for the 75% coverage matrices (Q10/ compared with background genomic differentiation
Q30, respectively). The ingroup branch has no support label as between species. The idea of genomic islands of diver-
C. rubiginosus was designated the outgroup to root the tree. gence was questioned because these genomic regions
Northern flicker RAD-seq relationships 7

were shown to also coincide with reduced genetic diver- Program #0801522 at the University of Kansas, and NSF
sity (Cruickshank & Hahn, 2014). However, recent work grants to JDM and RGM (DEB-1406989 and DEB-
in natural stickleback populations (Gasterosteus aculeatus 1241181). The COBRE Genome Sequencing Core Labo-
complex; Marques et al., 2016) has provided data indicat- ratory, funded by NIH award number P20GM103638, pro-
ing genomic islands of divergence without reduced back- vided laboratory facilities and services.
ground genetic diversity. Genomic islands may
alternatively develop due to differential selection in spe-
cies that could arise via adaptation to local conditions. Funding
The common hybridization between the auratus, cafer, National Science Foundation, Division of Environmental
and chrysoides morphological groups, the strong potential Biology [grant DEB-1241181, DEB-1406989].
for adaptive variation (e.g., chrysoides lives in deserts vs.
the forest habitat of other forms), and lack of clear genetic
signal here using hundreds of SNPs, warrants future inves-
tigation in more depth in both genomic and geographic Supplemental data
scales. Supplemental data for this article can be accessed here: http://dx.
Here, we provided the first, albeit limited, sequence doi.org/10.1080/14772000.2016.1238020.
data of the Cuban chrysocaulosus morphological group.
In the 231 bp sequence we obtained, two individuals
shared a haplotype and were one bp different from other Data accessibility
flicker individuals (Fig. 2). This low level of divergence
All Sanger sequence data is uploaded to NCBIs GenBank
suggests either very recent colonization of Cuba by C.
(Accession #s: KX698407-KX698421), and all Illumina
auratus, or sufficient levels of gene flow between Cuba
sequence data are uploaded to NCBIs Sequence Read
and the mainland U.S. to preclude isolation and subse-
Archive with BioProject ID PRJNA338861.
quent sequence divergence. However, the differentiation
between chrysocaulosus and mainland birds should be
interpreted with caution until more data become available.
Lastly, we provide the first genetic data of the mexica-
Declaration of interests
noides group, which appears distinct in both mtDNA and The authors report that they have no conflicts of interest.
nuclear DNA (Figs 2, 4). Although it appears that the mex- The authors alone are responsible for the content and writ-
icanoides individuals are distinct, gene flow into the cafer ing of the paper.
group appears likely, as evidenced by STRUCTURE
(Fig. 4) analyses. Because these cafer individuals were
sampled far from the southern limit of the taxon ORCID
(Table 1), genetic samples of the cafer group closer to Joseph D. Manthey http://orcid.org/0000-0003-2765-7611
Central America are necessary to identify the level of Robert G. Moyle http://orcid.org/0000-0001-6513-2344
gene flow between these groups in more proximal
locations.
References
Andersen, M. J., Nyari, A. S., Mason, I., Joseph, L., Dumbacher,
J. P., Filardi, C. E., & Moyle, R. G. (2014). Molecular sys-
Acknowledgements tematics of the worlds most polytypic bird: The Pachyce-
We would like to thank the state, provincial, and federal phala pectoralis/melanura (Aves: Pachycephalidae) species
complex. Zoological Journal of the Linnean Society, 170,
agencies and wildlife officers for their cooperation and 566588.
help obtaining permits that contributed to this research.  S., & Navarro-Sig
Arbelaez-Cortes, E., Nyari, A. uenza, A. G.
We also wish to thank the curators and collection manag- (2010). The differential effect of lowlands on the phylogeo-
ers at the institutions that provided tissue samples critical graphic pattern of a Mesoamerican montane species (Lepi-
for completion of this study. These include: John Klicka docolaptes affinis, Aves: Furnariidae). Molecular
Phylogenetics and Evolution, 57, 658668.
and Sharon Birks (University of Washington, Burke Bandelt, H. J., Forster, P., & R ohl, A. (1999). Median-joining
Museum of Natural History), and James V. Remsen and networks for inferring intraspecific phylogenies. Molecular
Donna Dittmann (Louisiana State University Museum of Biology and Evolution, 16, 3748.
Natural Science Collection of Genetic Resources). This Barber, B. R., & Klicka, J. (2010). Two pulses of diversification
study was partially funded through an American Museum across the Isthmus of Tehuantepec in a montane Mexican
bird fauna. Proceedings of the Royal Society of London B:
of Natural History Frank M. Chapman Memorial Fund Biological Sciences, 277, 26752681.
Grant, an American Ornithologists Union Student Barrera-Guzman, A. O., Mila, B., Sanchez-Gonzalez, L. A., &
Research Grant, through the NSF IGERT C-CHANGE Navarro-Sig uenza, A. G. (2012). Speciation in an avian
8 J. D. Manthey et al.

complex endemic to the mountains of Middle America Leache, A. D., Harris, R. B., Rannala, B., & Yang, Z. (2013).
(Ergaticus, Aves: Parulidae). Molecular Phylogenetics and The influence of gene flow on species tree estimation: A
Evolution, 62, 907920. simulation study. Systematic Biology, 63, 1730.
Catchen, J. M., Amores, A., Hohenlohe, P., Cresko, W., & Post- Manthey, J. D., Klicka, J., & Spellman, G. M. (2011). Cryptic
lethwait, J. H. (2011). Stacks: Building and genotyping loci diversity in a widespread North American songbird: Phylo-
de novo from short-read sequences. G3: Genes, Genomes, geography of the Brown Creeper (Certhia americana).
Genetics, 1, 171182. Molecular Phylogenetics and Evolution, 58, 502512.
Chesser, R. T., Banks, R. C., Burns, K. J., Cicero, C., Dunn, J. Manthey, J. D., & Moyle, R. G. (2015). Isolation by environment
L., Kratter, A. W., Rising, J. D. (2015). Fifty-sixth sup- in White-breasted Nuthatches (Sitta carolinensis) of the
plement to the American Ornithologists Union: Check-list Madrean Archipelago sky islands: A landscape genomics
of North American birds. The Auk, 132, 748764. approach. Molecular Ecology, 24, 36283638.
Chifman, J., & Kubatko, L. (2014). Quartet inference from SNP Marques, D. A., Lucek, K., Meier, J. I., Mwaiko, S., Wagner, C.
data under the coalescent model. Bioinformatics, 30, 3317 E., Excoffier, L., & Seehausen, O. (2016). Genomics of
3324. Rapid Incipient Speciation in Sympatric Threespine Stickle-
Cruickshank, T. E., & Hahn, M. W. (2014). Reanalysis suggests back. Public Library of Science Genetics, 12, e1005887.
that genomic islands of speciation are due to reduced McCarthy, E. M. (2006). Handbook of Avian Hybrids of the
diversity, not reduced gene flow. Molecular Ecology, 23, World. New York, NY: Oxford University Press.
31333157. McCormack, J. E., Maley, J. M., Hird, S. M., Derryberry, E. P.,
del Hoyo, J., Elliott, A., Sargatal, J., Christie, D.A., & de Juana, Graves, G. R., & Brumfield, R. T. (2012). Next-generation
E. (Eds.). (2016). Handbook of the Birds of the World Alive. sequencing reveals phylogeographic structure and a species
Barcelona: Lynx Edicions. Retrieved from http://www.hbw. tree for recent bird divergences. Molecular Phylogenetics
com/ (accesed10 May 2016). and Evolution, 62, 397406.
Decker, J. E., McKay, S. D., Rolf, M. M., Kim, J., Alcala, A. M., Mila, B., McCormack, J. E., Casta~ neda, G., Wayne, R. K., &
Sonstegard, T. S., Babar, M. E. (2014). Worldwide pat- Smith, T. B. (2007). Recent postglacial range expansion
terns of ancestry, divergence, and admixture in domesticated drives the rapid diversification of a songbird lineage in the
cattle. Public Library of Science Genetics, 10, e1004254. genus Junco. Proceedings of the Royal Society B: Biological
Dufort, M. J. (2016). An augmented supermatrix phylogeny of the Sciences, 274, 26532660.
avian family Picidae reveals uncertainty deep in the family Miller, M. R., Dunham, J. P., Amores, A., Cresko, W. A., & John-
tree. Molecular Phylogenetics and Evolution, 94, 313326. son, E. A. (2007). Rapid and cost-effective polymorphism
Ellegren, H., Smeds, L., Burri, R., Olason, P. I., Backstr om, N., identification and genotyping using restriction site associated
Kawakami, T., Uebbing, S. (2012). The genomic land- DNA (RAD) markers. Genome Research, 17, 240248.
scape of species divergence in Ficedula flycatchers. Nature, Monroe, B. L., Banks, R. C., Fitzpatrick, J. W., Howell, T. R.,
491, 756760. Johnson, N. K., Storer, R. W. (1995). Fortieth supplement
Evanno, G., Regnaut, S., & Goudet, J. (2005). Detecting the to the American Ornithologists Union: Check-list of North
number of clusters of individuals using the software American birds. The Auk, 112, 819830.
STRUCTURE: A simulation study. Molecular Ecology, 14, Moore, W. S., Graham, J. H., & Price, J. T. (1991). Mitochon-
26112620. drial DNA variation in the northern flicker (Colaptes aura-
Fletcher, S. D., & Moore, W. S. (1992). Further analysis of allo- tus, Aves). Molecular Biology and Evolution, 8, 327344.
zyme variation in the northern flicker, in comparison with Ornelas, J. F., Sosa, V., Soltis, D. E., Daza, J. M., Gonzalez, C.,
mitochondrial DNA variation. Condor, 94, 988991. Soltis, P. S., Ruiz-Sanchez, E. (2013). Comparative phy-
Grudzien, T. A., Moore, W. S., Cook, J. R., & Tagle, D. (1987). logeographic analyses illustrate the complex evolutionary
Genic population structure and gene flow in the Northern history of threatened cloud forests of northern Mesoamerica.
Flicker (Colaptes auratus) hybrid zone. The Auk, 104, 654 Public Library of Science One, 8, e56283.
664. Pickrell, J. K., & Pritchard, J. K. (2012). Inference of population
Guindon, S., & Gascuel, O. (2003). A simple, fast, and accurate splits and mixtures from genome-wide allele frequency data.
algorithm to estimate large phylogenies by maximum likeli- Public Library of Science Genetics, 8, e1002967.
hood. Systematic Biology, 52, 696704. Pritchard, J. K., Stephens, M., & Donnelly, P. (2000). Inference
Hackett, S. J. (1996). Molecular phylogenetics and biogeography of population structure using multilocus genotype data.
of tanagers in the genus Ramphocelus (Aves). Molecular Genetics, 155, 945959.
Phylogenetics and Evolution, 5, 368382. Short, L. L. (1982). Woodpeckers of the World. Greenville, DE:
Johnson, K. P., & Sorenson, M. D. (1998). Comparing molecular Delaware Museum of Natural History.
evolution in two mitochondrial protein coding genes (cyto- Swofford, D. L. (2003). PAUP. Phylogenetic analysis using
chrome b and ND2) in the dabbling ducks (Tribe: Anatini). parsimony ( and other methods). Version 4. Sunderland,
Molecular Phylogenetics and Evolution, 10, 8294. MA: Sinauer Associates.
Jombart, T., & Ahmed, I. (2011). Adegenet 1.3-1: New tools for Wiebe, K. L., & Bortolotti, G. R. (2001). Variation in colour
the analysis of genome-wide SNP data. Bioinformatics, 27, within a population of northern flickers: A new perspective
30703071. on an old hybrid zone. Canadian Journal of Zoology, 79,
Jombart, T., Devillard, S., & Balloux, F. (2010). Discriminant 10461052.
analysis of principal components: A new method for the Wiebe, K. L., & Bortolotti, G. R. (2002). Variation in caroten-
analysis of genetically structured populations. BioMed Cen- oid-based color in northern flickers in a hybrid zone. The
tral Genetics, 11, 94. Wilson Bulletin, 114, 393400.
Klicka, J., Spellman, G. M., Winker, K., Chua, V., & Smith, B. T. Wiley, E. O., & Mayden, R. L. (2000). The evolutionary species
(2011). A phylogeographic and population genetic analysis of concept. In R. Meier (Ed.), Species Concepts and Phyloge-
a widespread, sedentary North American bird: The Hairy netic Theory: A Debate (pp 7089). New York, USA:
Woodpecker (Picoides villosus). The Auk, 128, 346362. Columbia University Press.
Northern flicker RAD-seq relationships 9

Winkler, H., & Christie, D. A. (2002). Family Picidae (wood- Zink, R. M. (2010). Drawbacks with the use of microsatellites in
peckers). In D. del Hoyo, A. Elliot, & J. Sargatal (Eds.), phylogeography: The song sparrow Melospiza melodia as a
Handbook of the Birds of the World Volume 7 - Jacamars to case study. Journal of Avian Biology, 41, 17.
Woodpeckers (pp 296555). Barcelona, Spain: Lynx
Edicions. Associate Editor: Gary Voelker

Das könnte Ihnen auch gefallen