Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Molecular Phylogenetics and Evolution 102 (2016) 265–277

Contents lists available at ScienceDirect

Molecular Phylogenetics and Evolution


journal homepage: www.elsevier.com/locate/ympev

Review

Phylogeography of endemic Xantus’ hummingbird (Hylocharis xantusii)


shows a different history of vicariance in the Baja California Peninsula
Cristina González-Rubio a, Francisco J. García-De León a,⇑, Ricardo Rodríguez-Estrella b
a
Laboratorio de Genética para la Conservación, Centro de Investigaciones Biológicas del Noroeste (CIBNOR), Av. IPN #195, La Paz, BCS 23096, Mexico
b
Laboratorio de Análisis Espacial, Ecología y Conservación, Centro de Investigaciones Biológicas del Noroeste (CIBNOR), Av. IPN #195, La Paz, BCS 23096, Mexico

a r t i c l e i n f o a b s t r a c t

Article history: Studies of phylogeographic patterns provide insight into the processes driving lineage divergence in a
Received 26 August 2015 particular region. To identify the processes that caused phylogeographic breaks, it is necessary to use his-
Revised 24 May 2016 torical information and a set of appropriate molecular data to explain current patterns. To understand the
Accepted 30 May 2016
influence of geological or ecological processes on the phylogeography of the only species of hummingbird
Available online 31 May 2016
endemic to the Baja California Peninsula, Hylocharis xantusii, mitochondrial DNA sequences of three con-
catenated genes (Cyt-b, COI and ND2; 2297 bp in total) in 100 individuals were analyzed. The spatial
Keywords:
analyses of genetic variation showed phylogeographic structure consisting of a north, central and south
Hylocharis xantusii
Baja California Peninsula
regions. According to estimated divergence times, two vicariant events are supported, permanent sepa-
Phylogeographic breaks ration of the peninsula and formation of the Gulf of California at 5 mya and temporary isolation of the
Divergence times southern region at the Isthmus of La Paz at 3 mya. The temporal frame of genetic differentiation of
intraspecific haplotypes indicates that 90% of haplotypes diverged within the last 500,000 years, with a
population expansion 80,000 years ago. Only four haplotypes diverged 2.2 my and occurred in the south
(Hxan_36, 38 and 45), and north (Hxan_45 and 56) regions; only haplotype 45 is shared between south
and north populations. These regions also have the most recent haplotypes from 12,500 to 16,200 years
ago, and together with high levels of genetic diversity, we suggest two refuge areas, the Northern and
Southern regions. Our results indicate that the phylogeographic pattern first results from vicariance pro-
cesses, then is followed by historical and recent climate fluctuations that influenced conditions on the
peninsula, and it is also related to oases distribution. This study presents the first investigation of phylo-
geography of the peninsular’ endemic Xantus’ hummingbird.
Ó 2016 Elsevier Inc. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
2. Materials and methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
2.1. Area of study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
2.2. Sampling and laboratory procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
2.3. Phylogeographic patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
2.3.1. Estimating genetic diversity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
2.3.2. Phylogenetic analyses. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
2.3.3. Analysis of molecular variance (AMOVA) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
2.3.4. Spatial analysis of molecular variance (SAMOVA) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
2.3.5. Isolation by distance’s test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
2.4. Historical events . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
2.4.1. Estimating divergence times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
2.4.2. Historical demography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
2.4.3. Population dynamics through time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
2.5. Identifying refuge areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270

⇑ Corresponding author.
E-mail address: fgarciadl@cibnor.mx (F.J. García-De León).

http://dx.doi.org/10.1016/j.ympev.2016.05.039
1055-7903/Ó 2016 Elsevier Inc. All rights reserved.
266 C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277

3. Results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
3.1. Phylogeographic pattern for the Xantus’ hummingbird . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
3.1.1. Genetic diversity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
3.1.2. Phylogenetic analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
3.1.3. Spatial analysis of genetic variation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
3.2. Historical events: vicariance versus Pleistocene climate changes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
3.2.1. Temporal framework of genetic differentiation of the Xantus’ hummingbird . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
3.2.2. Demographic history . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
3.3. Identifying refuge areas of the Xantus’ hummingbird . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
4. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
4.1. Phylogenetic analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
4.1.1. Using mtDNA as a single locus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
4.2. Phylogeography of the Xantus’ hummingbird . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
4.2.1. Vicariant events versus Pleistocene climatic changes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
4.2.2. Historical demography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
4.3. Defining refuge areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
4.4. Conservation implications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
5. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
Appendix A. Supplementary material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275

1. Introduction endemic species (Grismer, 2000; Riddle et al., 2000b; Rojas-Soto


et al., 2003). Semi-arid and arid regions have been most sensitive
Phylogeography has proved successful in explaining how distri- to climate change, particularly during interglacial periods, when
butions of natural populations have been influenced by historical low rainfall leads to expansion of arid vegetation, decreases in
events, such as geological changes and climate fluctuations. For quality and quantity of groundwater, soil degradation, salinization,
example, some studies of mitochondrial DNA sequences of North and reduction in vegetation cover (Gutierrez-Elorza, 1998). Since
American vertebrates and plants have shown that rapid climate the start of the Holocene (11,500 years; Gibbard and Van
change during Pleistocene glaciations, together with topographic Kolfschoten, 2004), several animal populations have been
and ecological barriers (e.g. vicariance), were fundamental in gen- restricted to their most environmental stable areas (refuges). The
erating phylogeographic structures within species and populations southern part of the peninsula is regarded as a refuge for some
(Hewitt, 2000; Riddle et al., 2000a; Zink et al., 2001; Lemmon et al., birds, as for example, the California gnatcatcher (Polioptila califor-
2007; Scoble and Lowe, 2010; Smith et al., 2011). Three generaliza- nica; Zink et al., 2013, 2001; Zink, 2002), and the California quail
tions are formulated from these works: (1) taxa were isolated into (Callipepla californica; Zink, 2002). In this area, they show high
independent lineages (species, subspecies, or phylogroups), first genetic diversity compared with northern localities of the
because of major geological changes during the Pliocene epoch peninsula.
5.3–2.6 mya (Orndorff et al., 2007), (2) climatic shifts during the Xantus’ hummingbird Hylocharis xantusii (Lawrence, 1860)
Pleistocene about 2.6–1.8 mya (Polly et al., 2009), which con- belongs to the family Trochilidae, which has the highest degree
tributed to the isolation of populations in many areas and have of specialization mainly in morphological traits, related to flight
led to genetic divergence and speciation, and (3) highest genetic ability, physiology along with feeding adaptations, as well as
diversity in species or populations in areas that remained stable, behavior and sexual dimorphism, compared to all other families
the refuges during glacial times (Hewitt, 1996, 2000). of birds (Temeles et al., 2010; Berns and Adams, 2013). This
In the Baja California Peninsula (BCP), studies of phylogeo- medium-sized hummingbird displays a white post-ocular stripe
graphic patterns suggest dispersal and vicariance for many taxa contrasting greatly with a broad blackish auricular mask. Strong
(Riddle et al., 2000a,b; Zink, 2002; Zink et al., 2001, 2000; Crews sexual dimorphism is present: males have an orange bill and color-
and Hedin, 2006; Devitt, 2006; Riddle and Hafner, 2006a; Leaché ful plumage (Howell and Howell, 2000). The sister species, the
and Mulcahy, 2007; Leaché et al., 2009; Munguia-Vega, 2011). white-eared hummingbird Hylocharis leucotis (Hernández-Baños
Three major events explain the phylogeographic structure in the et al., 2014), is found in southern Arizona, Nicaragua, Guatemala
region: (1) Permanent separation of the peninsular and continental and in Mexico except the peninsula (Howell and Howell, 2000).
populations following the formation of the Gulf of California It’s distribution does not overlap with H. xantusii’ distribution. This,
(5 mya; Riddle et al., 2000a,b), (2) Temporary isolation of the pop- suggest that these two species diverged after the formation of the
ulation in the southern end of the peninsula (Cape Region) from peninsula (Zamudio-Beltrán, 2011). Little is known about the biol-
other peninsular populations following inundation at the Isthmus ogy, ecology, and distribution patterns of Xantus’ hummingbird,
of La Paz (3 mya; Riddle et al., 2000b), and (3) a Mid-peninsular but it is restricted to the central and southern parts of the penin-
seaway had temporarily isolated northern populations from south- sula with discontinuous distribution; it has two breeding seasons
ern populations (1 mya; Riddle et al., 2000a; Zink, 2002; Zink et al., each year, usually two chicks per nest, and is found from sea level
2001). This seaway break accounts for consistent separation of sev- to about 1800 m (Howell and Howell, 2000).
eral species (Munguia-Vega, 2011). Additionally to vicariant Three features endorse this hummingbird as a model to under-
events, changes in local environmental conditions (e.g. topography, stand the influence of geological and ecological dynamics in penin-
altitude) in conjunction with Pleistocene and Holocene climatic sular populations of birds, mammals, and reptiles. First, it is the
fluctuations (2.6 mya to the present) also influenced genetic struc- only neotropical hummingbird species endemic to the peninsula
turing in peninsular populations (Leaché and Mulcahy, 2007; (Arriaga et al., 1990); second, this species inhabits the Sierra La
Garrick et al., 2013, 2009). Laguna region (Arriaga et al., 1990), which is the most outstanding
The peninsula is a geologically dynamic region that includes an region in terms of the number of endemic taxa, and also the only
array of landscapes and habitats as well as a large number of area that contains pine-oak forests within the south of the
C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277 267

peninsula (Rodríguez-Estrella et al., 2005, 1999). Third, outside of 2014, 2007; Hernández-Baños et al., 2014). Species selection was
the Sierra La Laguna, it is mostly restricted to oases, refuges that based on the following criteria: H. leucotis and H. xantusii are sister
contain relict mesic habitats from the Pleistocene (Maya et al., species; C. latirostris occurs with H. leucotis; Calypte costae occurs
1997; Valero et al., 2014). The oases, dispersed throughout the with H. xantusii; Amazilia versicolor is related to Hylocharis species
peninsula, are highly valuable for the unique biotas that they and C. latirostris; and A. colubris is related to C. costae. The genomes
support, containing rich food sources for many small animals of the last two species were used as reference to delineate testing
(Arriaga et al., 1997; Rodríguez-Estrella et al., 2005, 1999, sequences. We downloaded complete genomes from the GenBank
1997). Understanding the phylogeographic pattern of the Xantus’ database (A. versicolor accession number KF624601, and A. colubris
hummingbird and its relation to historical processes, is particularly no. EF532935).
important in evolutionary and adaptive terms, since it is the only Samples were obtained from hummingbirds in the field (except
neotropical endemic species that has adapted to an arid peninsula A. versicolor and A. colubris). We collected samples of Xantus’ hum-
compared to other endemic species that have Nearctic origin mingbird along its entire geographic distribution; the locations and
which presently only inhabit the pine-oak and tropical deciduous sample numbers per site are shown in Fig. 1 and Table 1 respec-
forests in southern Baja California peninsula and have failed to tively. We have included two mainland locations for H. leucotis
extend beyond this habitat (Rodríguez-Estrella, 1987). and C. latirostris (Fig. 1). Samples of C. costae were obtained within
In this study, we analyzed Xantus’ hummingbird phylogeo- the peninsula’s locations. Birds were captured with mist nets, then
graphic structure, specifically the intraspecific evolutionary history blood was collected from the tarsus, and birds were released. No
in light of current biogeography and historical climatic conditions. injuries were noted during blood sampling; standard procedures
We analyzed fragments of three mitochondrial DNA concatenated were used for managing the hummingbirds.
genes of individuals covering its entire geographic distribution to Genomic DNA from the blood was obtained using the rapid salt
test whether: (1) there is a phylogeographic pattern of intraspecific extraction protocol (Aljanabi and Martínez, 1997). We analyzed
lineages, (2) phylogeographic patterns are related to specific his- three fragments of mtDNA in the six related hummingbirds’ spe-
torical events, either a vicariant event or Pleistocene climate cies: (1) nicotinamide adenine dinucleotide dehydrogenase sub-
changes, and (3) there are refuge areas representing Pleistocene unit 2 (ND2), which were amplified with H6313 and L5219
climatic conditions. primers (Sorenson et al., 1999); (2) cytochrome c oxidase subunit
1 (COI) amplified with primers H8191 and L7165 (De Filippis and
Moore, 2000); and (3) cytochrome b (Cyt-b) amplified with primers
2. Materials and methods
H16065 and L14841 (Kocher et al., 1989). For the hummingbird
field samples, all sequences were deposited in GenBank (http://
2.1. Area of study
www.ncbi.nlm.nih.gov/).
All PCR reactions were performed in 14 lL total volume, con-
The Baja California Peninsula contains a wide array of habitats.
taining about 40 ng of DNA, 1 PCR buffer (20 mM Tris-HCl,
Vegetation types include xerophytic scrub (including the arroyos’
50 mM KCl at pH 8.4), 0.2 mM of each dNTP, 0.4 lM of each primer,
vegetation), pine-oak forests (only in the Sierra La Laguna, which
2.0 lM MgCl2, and 0.5 unit Taq DNA polymerase (Invitrogen, Carls-
is above 1400 m), tropical deciduous forest (only in the Sierra La
bad, CA). The temperature profiles included initial denaturation at
Laguna, 300–800 m), coastal vegetation, mesic vegetation of oases
94 °C for 5 min, 30 cycles at 94 °C for 1 min each, 30 s at 55 °C
(including palms and reed grass), and similar vegetation of adja-
annealing temperature, and a final extension at 72 °C for 5 min.
cent off-coast islands in the Pacific and the Gulf of California. Mean
Sequences were reported by DNA sequencing services (Genewiz,
annual precipitation ranges from 100 mm in the Vizcaino Desert to
Plainfield, NJ; http://www.genewiz.com/; Macrogen (Seoul, South
more than 765 mm in the Sierra La Laguna (Rodríguez-Estrella
Korea; http://www.macrogen.com/eng/). Sequences were edited
et al., 2005). An important feature of the arid and semi-arid regions
with ChromasPro 1.7.6 and aligned with MUSCLE in MEGA 6
of the peninsula is the highly seasonal nature of rainfall, summer
(Tamura et al., 2013), and revised manually, using UltraEdit Profes-
tropical storms bring considerable flash floods, mostly in two-
sional 14.20.1.
thirds southern portion of the peninsula. In more arid areas, tem-
peratures can rise to 50 °C, while they do not exceed 30 °C in
2.3. Phylogeographic patterns
the humid Sierra La Laguna and other oases. Oases are separated
by at least 10 km, with some that are more than 50 km from the
In order to determine if there is a phylogeographic structure for
nearest neighbor, their size varies from 0.05 to 2.7 km2, and only
the Xantus’ hummingbird, we characterized the genetic diversity
48% of them have surface water. Oases contain the highest bird
by conducting three approaches: a phylogenetic analysis, molecu-
species richness, diversity and abundances from all the peninsula
lar variance and spatial distribution of molecular variance analysis,
habitats (Arriaga and Rodríguez-Estrella, 1997; Rodríguez-Estrella
and Isolation by distance’ test.
et al., 1997). H. xantusii, is a species related to oases regions, mainly
those with surface water and vegetation of arroyos. It is found
2.3.1. Estimating genetic diversity
more abundant in the oases of greater size, particularly towards
The characterization of genetic diversity at the intraspecific
the southern region of the peninsula. In this region, the species is
level was estimated from nucleotide and haplotype diversity, the
very common, especially after rains in the pine-oak forests of Sierra
number of segregating sites, and the number of haplotypes. These
La Laguna (Rodríguez-Estrella, 1997, 1987; Arriaga et al., 1990). It
analyses were performed for each group (see below), using sepa-
is found less abundant in the northern part of the species’ distribu-
rated mtDNA genes (Cyt-b, COI, ND2), and concatenated sequences,
tion, where oases are more distant from each other (Rodríguez-
all this was performed using DnaSP 5 software (Librado and Rozas,
Estrella, 1997).
2009).

2.2. Sampling and laboratory procedures 2.3.2. Phylogenetic analyses


First, we used maximum likelihood in GARLI 2.0 (Zwickl, 2011)
Six related hummingbirds’ species were used in this work: to recover phylogenetic relationships as well as to obtain a non-
Hylocharis xantusii, H. leucotis, Cynanthus latirostris, Amazilia versi- parametric bootstrap support value for the nodes. Individual solu-
color, Calypte costae, and Archilochus colubris (McGuire et al., tions of trees were selected after 10,000,000 generations and when
268 C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277

Fig. 1. Map of the Baja California Peninsula showing Xantus’ hummingbird (Hylocharis xantusii) localities sampled in the present study (labeled 1–15). The number of
individuals per locality is shown in Table 1. Three major vicariant events are showed in dashed lines (Gulf of California, Isthmus of La Paz, and Mid-peninsular seaway; Riddle
et al., 2000a,b; Zink et al., 2001; Zink, 2002). Mainland localities (16 and 17) belongs to samples of Hylocharis leucotis and Cynanthus latirostris used in this work.

the total improvement in likelihood score was 0.05. We also per-


formed 1000 bootstrap replicates, each included two independent
Table 1 runs. Second, phylogenetic reconstruction and posterior probabil-
Sampling sites used in this work. Species sample sizes are indicated for the endemic
Xantus’ hummingbird (Hylocharis xantusii), and for mainland species (H. leucotis and
ity estimates from MrBayes 3.2 were obtained using Bayesian
Cynanthus latirostris). Numbers of sampling sites correspond to the numbers in Fig. 1. inference (Ronquist et al., 2012). We used two parallel runs, one
‘cold’ and three ‘hot’ chains for 20,000,000 generations, sampling
Sampling sites Species sample size
every 10,000 generations until a standard deviation value less than
Hylocharis Hylocharis Cynanthus 0.01 of split frequencies was reached. Data was partitioned by the
xantusii leucotis latirostris
mtDNA gene, then, using the best-fit model for each one, it was
Peninsula analyzed in jModelTest 2 considering the corrected Akaike infor-
1 Santa Gertrudis 9
mation criteria (Darriba et al., 2012). We used a 25% burn-in, and
2 San Ignacio 5
3 San Zacarías 5 it was calculated a majority rule consensus tree. Each phylogenetic
4 San José de Magdalena 10 analysis included concatenated sequences (Cyt-b, COI, ND2) of five
5 San Nicolás 6 species as outgroup: Hylocharis leucotis (seven haplotypes, this
6 Carambuche 10 study), Cynanthus latirostris (one haplotype, this study), Amazilia
7 San Isidro 3
8 San José 4
versicolor (one haplotype, accession number KF624601), Calypte
9 San Miguel 8 costae (five haplotypes, this study), and A. colubris (one haplotype,
10 San Javier 4 accession number EF532935).
11 La Soledad 9
12 San Blas 3
13 San Dionisio 3 2.3.3. Analysis of molecular variance (AMOVA)
14 Sierra de la Laguna 11 There was no geographical criteria for delineating populations,
15 Santiago 9
so we performed AMOVA’s using Arlequin 3.5 (Excoffier and
Mainland Lischer, 2010), following these hypotheses (1) we considered two
16 Álamos, Sonora 6 4
17 El Fuerte, Sinaloa 4 3
groups, assuming that the vicariant Isthmus of La Paz separated
the south from other regions; (2) three groups, assuming the
C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277 269

Isthmus of La Paz, and that the vicariant Mid-peninsular seaway Cortés et al., 2014; Licona-Vera and Ornelas, 2014). We ran the
separated the north from central region; and (3) we considered analyses for 1  108 generations, sampling every 10,000 iterations,
11 groups, assuming that each oasis of the eleven oases supports using as prior the Yule speciation tree, a UPGMA starting tree, and
one phylogeographic lineage. These oases are: Santa Gertrudis, a log-normal relaxed uncorrelated molecular clock (Drummond
San Ignacio (including San Ignacio and San Zacarías), San José de and Rambaut, 2007). After the analysis, we used TreeAnnotator
Magdalena, San Nicolás, La Purísima (including Carambuche and 1.7.4 (Rambaut and Drummond, 2012) to generate a tree file com-
San Isidro), Los Comondús (including San José and San Miguel), piling the data, with a 25% burn-in and a posterior probability limit
San Javier, La Soledad, San Blas, Sierra La Laguna, and Santiago of 0.5. We used Tracer 1.6 (Rambaut and Drummond, 2009) to
(Fig. 1). assess chain convergence and determine that our sample of the
posterior distribution of the likelihoods had reached a sufficiently
2.3.4. Spatial analysis of molecular variance (SAMOVA) effective sample size (ESS), and therefore, stationarity.
We used SAMOVA 1.0 (Dupanloup et al., 2002) and K2P dis-
tances to detect major genetic discontinuities among sampling 2.4.2. Historical demography
sites, explicitly to test phylogeographic structure in H. xantusii. This Historical demography of the Xantus’ hummingbird was
method maximizes the proportion of genetic variance; based on assessed by separating haplotypes into the most significant num-
the differences found within the number of groups defined by ber of populations resulting from SAMOVA analyses. Each popula-
the user (K), assigning localities to populations, and with the tion was analyzed independently. We sought evidence for sudden
assumption that they must be geographically adjacent and genet- population growth in each population, for this we constructed mis-
ically homogeneous (Dupanloup et al., 2002). match distributions of pairwise nucleotide differences between
In order to find the optimal K value in which FCT (i.e., genetic haplotypes, using Arlequin 3.5 (Excoffier and Lischer, 2010). Mis-
variance owing to divergences between groups) was the highest match distributions are usually ragged in populations at demo-
and most significant, we tested different K values. We established graphic equilibrium, but are unimodal in populations that have
the hypotheses for AMOVA: K = 2, vicariant event of Isthmus of La experienced recent expansion (Rogers and Harpending, 1992).
Paz; K = 3, vicariant event of Mid-peninsular seaway, three geo- We estimated the raggedness index (Harpending, 1994), which in
graphic regions; and K = 11, oases sampled for H. xantusii. To verify the presence of relatively stable populations shows large values
consistency, we performed the analysis five times for each K value, and exhibits a multimodal distribution. Parameters used in the
with statistical significance of 1000 permutations. spatial expansion model were estimated from the data, assuming
h = 1, population size after expansion, as infinite. Goodness of fit
2.3.5. Isolation by distance’s test in spatial expansion was assessed by calculating the significance
We applied the model of Isolation by distance (Wright, 1943) in of the raggedness index through 1000 simulations that estimated
order to understand whether it exist a correlation between geo- the probability of obtaining an index less than the observed value
graphical and genetic distances between different pairs of Xantus’ (Harpending, 1994).
hummingbird populations. We carried out the analysis using the Additional evidence to support population expansion was
Isolation by Distance web service (http://ibdws.sdsu.edu/; Jensen obtained by applying the neutrality test by the D statistic
et al., 2005), with the same grouping hypotheses described in (Tajima, 1989). In this test, the significant positive numbers belong
AMOVA section. to a population that has recently undergone a bottleneck, whereas
significant negative numbers belong to a population that has
2.4. Historical events recently begun an expansion. An insignificant number is consistent
with a population at drift-mutation equilibrium. Population expan-
Phylogeographic patterns related to historical events, either a sion was also assessed with Fs (Fu, 1997). A negative number of Fs
geological event or Pleistocene climate change, were tested by per- evidence an excess number of genetic variants, as it would be
forming divergence time estimates which allowed establishing a expected from recent population expansion or from genetic hitch-
time frame for genetic differentiation and the historical demogra- hiking, while a positive number of Fs evidence a deficiency in
phy of Xantus’ hummingbird. genetic variants, as it would be expected from a recent population
bottleneck or in the presence of population structure (Fu, 1997). All
2.4.1. Estimating divergence times analyses were performed with Arlequin 3.5.
We used BEAST 1.8 (Drummond et al., 2012) to estimate the We analyzed the relationship among sampled haplotypes, in
time of divergence of the Xantus’ hummingbird and phylogenetic the program Network 4.6.1.2 performing a network haplotype
analysis, with three calibration points obtained from fossils and reconstruction through the median-joining algorithm (Bandelt
biogeographical events: (a) the external calibration point was 12 et al., 1999; http://www.fluxus-engineering.com/).
my, which is an estimated divergence time using fossil evidence,
that separates the Central and North American regions from the 2.4.3. Population dynamics through time
South American region (Bleiweiss, 1998), (b) 10 my, which approx- Past population dynamics for the Xantus’ hummingbird were
imately represents the formation of two major volcanic arcs, the examined in a coalescence-based framework using Extended Baye-
Sierra Madre Occidental and the Trans-Mexican volcanic belt sian Skyline Plot approach in BEAST (EBSP; Heled and Drummond,
(Ferrari et al., 2000), events that separate the northern Mesoamer- 2008). The skyline plot is a piecewise-constant model of popula-
ican hummingbird species (bees) from the north-central and tion size that fits a range of demographic scenarios. It has proved
northern American hummingbird species (emeralds), and (c) 5 very useful as a model selection tool used to indicate the most
my, approximate age of separation of the peninsula from mainland appropriate demographic model for any given data set. The EBSP
Mexico (Riddle et al., 2000a,b), separating the peninsular species model uses standard Markov chain Monte Carlo (MCMC) sampling
(Hylocharis xantusii) from the emerald group. The intraspecific sub- procedures. This allows estimating a posterior distribution of effec-
stitution rate of H. xantusii was calibrated according to rates previ- tive population size through time directly from a sample of gene
ously reported for those genes in birds; we used 2.07  10 8 sequences, given any specified nucleotide-substitution model. This
substitutions/site/year for Cyt-b; 1.6  10 8 substitutions/site/year method includes credibility intervals for the estimated effective
for COI, and 2.9  10 8 substitutions/site/year for ND2, (Weir and population size at every point in time, back to the most recent
Schluter, 2008; Qu and Lei, 2009; Sly et al., 2011; Arbeláez- common ancestor of the gene sequences (Drummond et al.,
270 C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277

2005). Analyses were run using a strict molecular clock, and the 1993), with some invariable and variable sites, following a discrete
HKY + G + I substitution model. The analysis was performed using gamma distribution with four categories (TrN + I + G).
separated mtDNA genes, and concatenated sequences (three times Based on the analysis of 56 Xantus’ haplotypes defined by con-
using the three mutation rates (1.6%, 2.1%, and 2.9% per site per catenated sequences, both maximum likelihood and Bayesian phy-
my) for 50 million iterations. Convergence, stationarity, effective logenetic reconstructions showed the same topology (Appendix A),
sample size for each parameter, and appropriate burn-in, were and had high statistical support; a main clade included Hylocharis
evaluated in the analysis using Tracer. xantusii, H. leucotis, Cynanthus latirostris, and Amazilia versicolor
(emerald clade). Within the emerald clade, all Xantus’ humming-
2.5. Identifying refuge areas bird haplotypes form a well supported (100%) clade by both phylo-
genetic methods. However, a clear phylogeographic pattern was
Refuge areas are expected to be restricted geographical areas not observed, as nine (Hxan_01, 04, 05, 07, 45, 09, 16, 17, 03) of
during Pleistocene climatic fluctuations, where populations main- the 56 haplotypes identified in Xantus’ hummingbird are shared
tained their genetic diversity; if individuals are restricted to among distant oases. All Xantus’ haplotypes assigned to each indi-
refuges for a long time it would be expected a genetic structure vidual are listed in Appendix B.
of isolated populations (e.g. Reding et al., 2012; Zink et al., 2013). Haplotype 01 was found at San Blas and San Javier (340 km
When conditions are restored, a process of dispersion could pro- apart) oases; haplotypes 04 and 05 were found at Santiago and
duce a gradient of genetic diversity or isolation by distance pattern San Ignacio (700 km); haplotype 07 was found at Santa Gertrudis
with or without genetic flow (e.g. Hull and Girman, 2004; González and La Soledad (600 km); haplotype 45 at Santa Gertrudis and
et al., 2011). We analyzed the existence of possible areas based on Sierra La Laguna (800 km); haplotype 09 at Carambuche and San
high levels of genetic diversity, and presence of ancestral and Dionisio (450 km); haplotype 16 at San Dionisio and San Nicolás
recent haplotypes as evidences of long-term refuge areas related (475 km), and is one of the ancestral haplotypes; haplotype 17 at
to climatic fluctuations. San Ignacio and San Isidro (220 km); and haplotype 03 at Santiago
and San Isidro (485 km).
3. Results
3.1.3. Spatial analysis of genetic variation
3.1. Phylogeographic pattern for the Xantus’ hummingbird AMOVA analyzes showed no significant differences for defining
genetic structure. However, when we incorporate geographic
3.1.1. Genetic diversity information to SAMOVA and isolation by distance analyzes, it
We sequenced 100 hummingbirds for Cyt-b (811 bp), COI was possible to determine a genetic structure related to south, cen-
(804 bp), ND2 (682 bp), and concatenated mtDNA (2297 bp). The tral and north regions, as well as relative to the 11 oases distribu-
lowest diversity value were obtained for Cyt-b gene, observing tion. Variations among groups revealed significant differences
11 haplotypes, in which haplotype diversity = 0.6851, nucleotide (Table 3), with the highest percentage by K = 3 (39.5%,
diversity = 0.0012, and mean number of pairwise differences = FCT = 0.3956, p < 0.05); south region included sites 11–15; central
0.924. The greatest number of haplotypes and haplotype diversity regions included sites 6–10, while north region included sites
were in COI gene (27 and 0.8811, respectively). In nucleotide 1–5 (Fig. 1). Isolation by distance analysis (Fig. 2), only showed a
diversity and mean number of pairwise differences, the highest significant correlation between geographical and genetic distances
numbers occurred in ND2 gene (0.0097 and 4.107, respectively). between different oases (K = 11; r = 0.1769, p = 0.002).
In concatenated sequences, we identified 56 unique haplotypes
Table 3
at 39 polymorphic sites; haplotype diversity was 0.9682, nucleo- SAMOVA results and percentage of variation depending of K defined groups.
tide diversity was 0.0048, and mean number of pairwise differ-
ences was 6.529. Genetic diversity for each region determined by Source of variation K=3 K = 11
% variation % variation
spatial analyses (see below) were similar, with the exception of (fixation index) (fixation index)
north population, with the lowest haplotype diversity (0.908)
Among groups 39.57 (0.3956**) 31.73 (0.3172***)
and mean number of pairwise differences (6.12) (Table 2).
Among populations 5.98 ( 0.0989 ns) 27.51 ( 0.4029 ns)
within groups
3.1.2. Phylogenetic analysis Within populations 66.41 (0.3358 ns) 95.78 (0.0422 ns)
The best-fit DNA evolution model for all mtDNA genes was the ns = non significant.
Tamura-Nei model, which allow variable base frequencies, equal **
p < 0.05.
***
transversion rates, and variable transition rates (Tamura and Nei, p < 0.001.

Table 2
Summary of genetic diversity and demographic statistics of the Xantus’ hummingbird endemic of the BCP, based on mtDNA haplotypes (Cyt-b, COI, ND2, concatenated
sequences).

Groups n S H h p Mean no. pairwise differences Raggedness index D’Tajima (ns) Fu’s Fs
Cyt-b (811 bp) 100 8 11 0.6851 0.0012 0.924 0.1249 ns 1.0876 6.197*
COI (804 bp) 100 15 27 0.8811 0.0033 2.442 0.0242 0.4396 17.786**
ND2 (682 bp) 100 17 19 0.7846 0.0097 4.107 0.0439 ns 0.709 2.665*
Concatenated (2297 pb) 100 39 56 0.9682 0.0048 6.529 0.0084 ns 0.2698 45.004**
South 35 29 27 0.9802 0.0063 8.498 0.0109 ns 0.3094 7.417**
Central 29 35 21 0.9872 0.0093 12.542 0.0421 ns 0.7884 3.050*
North 35 24 23 0.9077 0.0045 6.123 0.0366 ns 0.0964 5.298*

ns = non significant.
n = sample size; S = segregating sites; H = number of haplotypes; h = haplotype diversity; p = nucleotide diversity.
South, north and central = regions identified by SAMOVA results and confirmed by Isolation by distance analyses.
*
p < 0.05.
**
p < 0.001.
C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277 271

3.2. Historical events: vicariance versus Pleistocene climate changes

3.2.1. Temporal framework of genetic differentiation of the Xantus’


hummingbird
BEAST results suggest that H. xantusii diverged from its sister
species (H. leucotis) during the Pliocene, 3.6 mya (Appendices C
and D), after the formation and separation of the peninsula from
the mainland (about 5 mya). Sequences belonging to Calypte costae,
a co-distribute species with H. xantusii, suggest divergence at 3.5
my after the separation of the peninsula.
For genetic differentiation of Xantus’ hummingbird haplotypes,
the BEAST analysis suggests separation among groups of haplo-
types during the late Pliocene and early Pleistocene, 3–2 mya
(Fig. 3); however, for almost 90% of the haplotypes, divergence
started within the last 500,000 years, and approximately 40% of
the haplotypes are less than 100,000 years old.
Three main Xantus’ haplotype groups coincide with the three
geographic regions, with few shared haplotypes between
populations.
In the southern region, the time to the most recent common
Fig. 2. Isolation by distance analysis showing a significant correlation between ancestor (tMRCA) is 2.14 my, which is the age of one of the ances-
geographical and genetic distances between different oases (r = 0.4769, p = 0.002). tral haplotypes occurring in the Sierra La Laguna area. The more
recent haplotypes are about 16,200 years old and were found at

Fig. 3. Divergence time estimates in years (HPD and 95% confidence interval) for the Xantus’ hummingbird, based on mtDNA concatenated sequences (Cyt-b, COI, ND2), using
log-normal relaxed uncorrelated molecular clock and UPGMA starting tree implementing in BEAST (Drummond et al., 2012). Calibration points were: (a) ancestral divergence
time estimation using fossil calibration (12 mya; Bleiweiss, 1998), (b) formation of volcanic arcs (10 mya; Ferrari et al., 2000), and (c) complete separation of the BCP (5 mya;
Riddle et al., 2000a,b). The gray dotted line indicates the divergence time for most of the haplotypes (500,000 years ago). Only ancestral and recent haplotype are show as
terminal tips including their divergence time and location in parenthesis.
272 C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277

Santiago, and also in central region (San José and San Javier sites). 3.2.2. Demographic history
Haplotypes of La Soledad (site above the Isthmus of La Paz) began Mismatch distribution of pairwise differences among
to diverge 460,000 years ago. The southern region shows evidence haplotype-concatenated sequences for each region contain marked
that it began to diverge after the formation of the water channel bimodal curves (Fig. 4), contrasting with expected unimodal pat-
formed in the Isthmus of La Paz (3 mya), this indicates that this tern in populations that experienced recent demographic expan-
vicariant event served as a barrier to isolated the Cape region. sion. Demographic parameters, such a Harpending’s Raggedness
In the Central region, the tMRCA is 922,000 years old. The oldest index (Table 2) indicated that the observed mismatch distribution
haplotype in this region is 780,000 years, and it is found in Caram- did not differ significantly from the distribution expected under
buche, but it is also found in north region (San José de Magada- population expansion. On the other hand, neutrality test showed
lena). The second oldest has 480,000 year and it is found at San differing results: Tajima’s D statistic did not show significant devi-
José. ations of strict neutrality in all regions (Table 2), suggesting demo-
The Northern region (particularly San Nicolás oasis) contains graphic stability; however, Fu’s FS statistic showed negative and
the most ancestral haplotype, 2.5 my, which also was found at significant differences, suggesting an excess of low frequency hap-
the Sierra La Laguna (southern region); the second ancestral haplo- lotypes, which is expected if H. xantusii has experienced recent
types have 2.06 my and they were found at San Nicolás and Sierra demographic expansion.
La Laguna. For the other haplotypes in this region, the tMRCA is 1.6 The results of Median-joining network for each individual gene
my; and the oldest haplotype is 879,500 years, it was found at (Appendix E) show a complex pattern on the relationships between
Santa Gertrudis; this region also includes the most recent haplo- haplotypes. A network of two haplotypes with high frequency
types, 12,500 years old, which also were found in the southern interconnects with other lower frequency, each forming a star con-
region. figuration. This strongly suggests that, in each network, the two

Fig. 4. Mismatch distribution (left), and Extended Bayesian skyline plots (right; dashed lines indicate 95% HPD intervals) are displayed for each region for the Xantus’
hummingbird, using mtDNA concatenated genes (Cyt-b, COI, ND2). In skyline plots the south and north region underwent and expansion beginning 80,000 years ago,
whereas central region expanded to a lesser degree 1,000,000 years ago.
C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277 273

groups of haplotypes evolved independently for different time Incomplete lineage sorting can also lead to pronounced differ-
periods. This haplotype spatial pattern is also consistent with the ences between geographic distribution and mtDNA or nuclear
bimodal mismatch distribution (Fig. 4). This pattern remains, but DNA variation. Following the formation of two daughter species
with more complexity for sequences concatenated (Appendix E). from an ancestor, lineage sorting is more rapid in the case of
The EBSP analyses, based on mtDNA concatenated genes (Fig. 4) mtDNA due to its lower effective population size (Funk and
were similar to each other when using different mutation rates, Omland, 2003). Therefore, incomplete lineage sorting in young
suggesting that the south and north regions underwent an species can potentially result in differences in population struc-
expansion process that began 80,000 years ago. Central region tures inferred from the two kinds of markers. Sex-biased dispersal
expanded to a lesser degree, 1,000,000 years ago. Because the tim- may also lead to such differences (Galtier et al., 2009).
ing for the expansion are dependent on the calibration data used, The mitDNA genes used in this work (Cyt-b, COI, and ND2) pro-
we do not interpret dates precisely, but only found that there was vide sufficient resolution for identifying intraspecific differentia-
at least one expansion process in the central region, and potentially, tion in some endemic birds (Arbeláez-Cortés et al., 2014). Our
a more recent one in the southern and northern regions, where we phylogenetic analyses supported the relationships between Xan-
found that there is a greater number of unique haplotypes. tus’ hummingbird and their close relatives (McGuire et al., 2014,
2007; Hernández-Baños et al., 2014), suggesting that the genetic
3.3. Identifying refuge areas of the Xantus’ hummingbird resolution in our study was robust for describing the interspecific
genetic variation of the six species examined in our study (Appen-
Based on genetic diversity (high haplotype diversity; Table 2 dix A).
and Appendix F) and presence of ancestral haplotypes (Fig. 3),
there are two main refuge areas suggested, the northern and 4.2. Phylogeography of the Xantus’ hummingbird
southern regions, which are also regions of diversification within
the species, based on the presence of recent haplotypes (Fig. 3). Analyses of intraspecific lineages of H. xantusii indicate neither a
well-resolved topology tree nor a structure that was completely
concordant with geographic distribution. Some authors suggest
4. Discussion that trees may appear conflicting if they have unequal numbers
of terminal, which can result from a lack of differentiation in
The deserts in the Baja California peninsula act as areas of ende- response to a barrier or because some lineages have experienced
mism, resulting in some clear patterns of diversification that are extinctions (Zink et al., 2000). For many species adapted to xeric
mainly explained by vicariant events (Riddle et al., 2000a, 2000b; conditions, such as the birds of the Baja California Peninsula, which
Zink, 2002). Although regional differentiation of birds of the penin- originated later than their relatives on the mainland, climatic
sula has been explained by vicariance, no clear phylogeographic changes of the Pleistocene erased biogeographical records in phy-
divisions had been observed as expected for highly vagile birds logenies (Zink et al., 2013, 2001; Zink, 2002). This seems to be the
(Zink et al., 2013, 2000), yet it has been described there is high case for the phylogenetic pattern of the Xantus’ hummingbird,
genetic diversity and ancestral haplotypes in the southern region since 90% of haplotypes diverged during the last 500,000 years.
(Zink et al., 2013, 2001; Zink, 2002). This work is the first example Furthermore, endemic species with a restricted distribution and
of phylogeographic structure in an endemic species of bird, Xantus’ high genetic variation, such as H. xantusii which has high levels
hummingbird, in the Baja California Peninsula. Our work showed a of haplotype and nucleotide diversity compared to other hum-
phylogeographic pattern emerged explained by vicariant events mingbirds and other birds in general (for details see Appendix F),
followed by climatic changes during the Pleistocene. retain high levels of mitochondrial genetic variation in local areas,
and this could represent early stages of the speciation process
4.1. Phylogenetic analyses (González et al., 2011).
SAMOVA analysis defined three genetically homogeneous and
4.1.1. Using mtDNA as a single locus geographically significantly different groups (south, central and
Patterns of divergence and relationships inferred from different north regions), but also a significant correlation between genetic
types of data (e.g. mtDNA versus nuclear DNA) can disagree (e.g. and geographic distances was observed (Fig. 2). Together with
Crews and Hedin, 2006), reflecting dissimilarities in modes of divergence time estimates, we suggested three ancestral popula-
inheritance or relative rates of evolution (Avise, 2000). Mitochon- tions occurred during Pliocene and Pleistocene: south, central,
drial DNA in animals evolves faster than nuclear DNA, and depend- and north (Fig. 3). Actually the oases and the Sierra La Laguna share
ing of the event that has caused divergence and speciation, the ecological features of mesic habitats that seem to be environments
ability to detect this event depends upon rates of evolution of these more stable than the rest of the arid peninsula, therefore serving as
types of DNA (Avise, 2000). For example, if vicariance has caused contemporary refuges for this hummingbird and other peninsular
speciation, mitochondrial genes may be the only partition species (Grismer and McGuire, 1993). In fact, the Xantus’ hum-
expected to reveal divergence due its faster substitution rate mingbird is more abundant in the Sierra La Laguna and oases in
(Avise, 2000). However, mitochondrial history may not reflect the the Sierra de la Giganta than in other areas, mainly sites where
true history of species because mtDNA gene variation may be water ponds are present in arroyos (Rodríguez-Estrella et al.,
caused by introgression (Funk and Omland, 2003). Mitochondrial 2005; Arriaga et al., 1990). For many endemic birds, phylogeo-
genes are less constrained by linkage to selected loci and hence graphic patterns suggest that at least some common factors were
are expected to introgress deeper than their nuclear counterparts extremely influential at the regional level; therefore, phylogeo-
(Funk and Omland, 2003). Mitochondrial introgression has been graphic patterns could depend more on climate changes than geo-
reported in hummingbird species (e.g. Rodríguez-Gómez and logical events to link species behavior and ecology (e.g. phylopatry,
Ornelas, 2015; Jiménez and Ornelas, 2016); however, introgression habitat, foraging niche, sexual selection), which are important fac-
of maternally inherited mtDNA should be unlikely to happen for tors relevant for species of recent origin with high genetic diversity
the Xantus’ hummingbird because all phylogenetic trees using sep- (Qu and Lei, 2009; González et al., 2011; Qu et al., 2011; Lara et al.,
arated mtDNA genes and concatenate sequences, showed all spe- 2012; Arbeláez-Cortés et al., 2014; Licona-Vera and Ornelas, 2014).
cies’ haplotypes in a single 100% supported clade, separating H. The distance between oases along the mid-peninsula may also be
xantusii from the other five species. influencing the current distribution pattern of the species, generat-
274 C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277

ing a degree of genetic structure at a local level, which may not be in shaping spatial patterns of intraspecific diversity of vertebrate,
totally defined using mitochondrial genes. In this sense, the use of plants, and invertebrate species on the peninsula (Crews and
genetic markers with higher mutation rate (e.g. microsatellite loci) Hedin, 2006; Garrick et al., 2013, 2009; Zink et al., 2013).
will define changes in population structure at a finer temporal-
space scale. 4.3. Defining refuge areas

4.2.1. Vicariant events versus Pleistocene climatic changes North American deserts provide a useful context to explore
Contemporary spatial patterns of birds and other vertebrates methods in phylogeography (Zink et al., 2013, 2001; Zink, 2002;
cannot be understood without knowing how species responded Devitt, 2006; Garrick et al., 2013, 2009). Many phylogeographic
to the geologic and climate history of their regions. For example, patterns of species in an arid region are a consequence of the arid-
the vicariant event of separation of the peninsula (5 mya) was an ification during Pleistocene interglacial periods, after which spe-
important factor that promoted genetic differentiation for H. xan- cies expanded from isolated refuge (Hull and Girman, 2004;
tusii and for other peninsular vertebrates (Riddle et al., 2000a,b; Mapelli et al., 2010; Reding et al., 2012; Zink et al., 2013, 2000).
Grismer, 2000). The time of separation of H. xantusii from its main- Pleistocene environments must have permitted continued differ-
land sister species (H. leucotis), began after the separation of the entiation of populations whose separations often had been initi-
peninsula (3.6 mya), which was also confirmed for its co- ated earlier (Avise and Walker, 1998). Sequences of mtDNA
distributed species, Calypte costae (3.5 mya). The second vicariant concatenate genes show that the ancestral haplotypes are
event supported in our study was the seaway at the Isthmus of restricted to the northern and southern regions, and are less wide-
La Paz (3 mya; Riddle et al., 2000b), where the Xantus’ humming- spread than derived haplotypes and likely are centered on the ori-
bird in the southern region has the tMRCA at 2.1 mya, confirming gins of expansion, suggesting potential refuge areas. This has been
that this seaway isolated the region and promoted genetic differ- suggested in other birds of the peninsula: The California gnat-
entiation from the central and northern regions. Therefore, the catcher (Polioptila californica; Zink et al., 2013, 2001; Zink, 2002)
phylogeographic history for the Xantus’ hummingbird is related and the California quail (Callipepla californica; Zink, 2002), but only
primarily to the historical vicariant event of the seaway in the Isth- in the southern region. Our study shows that southern and north-
mus of La Paz; secondly and more recently, related with climate ern regions with populations of the Xantus’ hummingbird were
change at the end of the Pleistocene, observing that the 90% of hap- refuges (ancestral refuges), and therefore deserve special attention
lotypes diverged during the last 500,000 years. because they support birds with high genetic variation that
includes all ancestral haplotypes.
4.2.2. Historical demography Several studies of Neotropical birds document phylogeographic
Evidence of population expansion is provided by Fu’s FS test, structure for species with discontinuous ranges resulting from
suggesting a significant excess of low frequency haplotypes in H. topographical barriers that restricted gene flow and drove genetic
xantusii in the three populations. However, differences in results divergence (González et al., 2011; Arbeláez-Cortés et al., 2014). In
of Fu’s and Tajima’s neutrality tests can be related to their statisti- these cases, the explanation of phylogeographic patterns is rela-
cal properties. Fu’s test is more powerful in detecting demographic tively straightforward because it is concordant with topographical
changes and genetic hitchhiking (Fu, 1997). heterogeneity. The Xantus’ hummingbird does not appear to be
Another indication of recent population expansion is provided tied to distributions with clear geographical barriers; however, this
by the results of the divergence time estimates and EBSP analysis, peninsula includes a heterogeneous array of landscapes, with arid,
which suggested a recent expansion, where 90% of haplotypes warm, and temperate regions (Grismer, 2000). The peninsula has
diverged in the last 500,000 years (Fig. 3), and more recently, a distinct desert subregions, the Desierto del Vizcaíno in the north
population expansion that began 80,000 years ago (Fig. 4). and the Desierto de Magdalena in the south (Grismer, 2000) that
The presence of the same haplotypes found in distant oases are separated by volcanic mountain ranges (the Sierra San Fran-
could reflect historical connectivity because some of them are cisco and Volcán las Tres Vírgenes). Extensive lava flows over the
ancestral haplotypes at low frequency, or incomplete linage sort- past 1.2 my and high aridity produced low-quality habitats. Thus,
ing. If migration is interrupted for several generations, bottleneck these arid areas could represent a topographical barrier to disper-
events could result in the elimination of haplotypes in local popu- sal and gene flow in the Xantus’ hummingbird. Northern and
lations. In this sense, strong genetic drift after reduction might southern regions should be recognized as long-term refuge, not
explain the occurrence of haplotype divergence at some sites just because the hummingbird in the Sierra La Laguna and San
(Avise and Walker, 1998). We would expect that pairs of individu- Nicolás have ancestral haplotypes, but these oases are a source of
als taken at random from the population differ, on average, by high haplotype diversity and many unique haplotypes.
many nucleotide substitutions, causing a mismatch distribution A relationship seems to exist between the major climatic
with two or more modes at large values of pairwise differences changes of the Late Pleistocene and Holocene on the peninsula
(Rogers and Harpending, 1992). Divergent haplotypes within the and the phylogeographic pattern and historical changes in demog-
same population can occur in species that have undergone severe raphy, inferred for the genetics of the Xantus’ hummingbird. Given
bottleneck processes; however, the presence of haplotypes with these associations, we suggest that the current oases wetlands and
low pairwise differences in bimodal mismatch distribution also their natural resources are the key habitat factor responsible for
suggests a population that experienced a recent expansion the current phylogeographic pattern of the Xantus’ hummingbird.
(Rogers and Harpending, 1992). This is highly important because This suggestion is also supported with the IBD analysis (Fig. 2),
a population with a complex demographic history, including both resulting a significant correlation only between geographical and
ancient population bottleneck and recent expansion, with the pres- genetic distances among the eleven oases. Thus, the phylogeogra-
ence of ancestral haplotypes, could promote retention of novel phy of the endemic Xantus’ hummingbird shows a different history
mutations and speciation processes (Avise and Walker, 1998; of vicariance for the fauna in the Baja California peninsula.
Scoble and Lowe, 2010). Our results point to a scenario where
the genetic composition of the Xantus’ hummingbird could be a 4.4. Conservation implications
result of a recent diversification process. Phylogeographical studies
in species in arid environments provide evidence that range expan- Phylogeographic analyses and the time frame of genetic differ-
sions during the Late Pleistocene and Holocene played a major role entiation of Hylocharis xantusii suggest two ancestral populations
C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277 275

that began to spread in the Late Pleistocene and early Holocene, tion of methods that improved the manuscript. The Secretaría de
promoting genetic differences between distant oases. Although Medio Ambiente y Recursos Naturales (SEMARNAT) issued the col-
this species show high genetic diversity at regional level, probably lecting permit (SGPA/DGVS/10182/11, 06983/13). Funding was
due to the high environmental, geological, climatic, and ecological provided by the Consejo Nacional de Ciencia y Tecnología (CONA-
heterogeneity, populations and their genetic diversity may be CYT project CB-2008-01-106925) granted to F.J.G.L. and (CONACYT
threatened (Grismer, 2000; Riddle et al., 2000b; Rojas-Soto et al., project 155956) granted to R.R.E. C.G.R.S. is a recipient of a fellow-
2003). In this sense, incorporating information from nuclear DNA ship (CONACYT 8492).
sequences (e.g. microsatellites) will allow a more comprehensive
history of the contemporary genetic structure, and together with
the information generated in this work, will help to establish better Appendix A. Supplementary material
conservation plans. Conservation of areas should include ancestral
populations as independent units with their environmental char- Supplementary data associated with this article can be found, in
acteristics. It has been observed that only one ecological variable, the online version, at http://dx.doi.org/10.1016/j.ympev.2016.05.
such as habitat (e.g. oases in the peninsula), can explain much of 039.
the variation in diversification rate in birds, reptiles and terrestrial
mammals (Wiens, 2015), and it should be considered at local-scale References
for explaining patterns of genetic differentiation.
For the Xantus’ hummingbird, refuge areas contain high genetic Aljanabi, S.M., Martínez, I., 1997. Universal and rapid salt-extraction of high quality
diversity and retention of ancestral variation. Thus, conservation of genomic DNA for PCR-based techniques. Nucleic Acids Res. 25, 4692–4693.
Arbeláez-Cortés, E., Milá, B., Navarro-Siguenza, A.G., 2014. Multilocus analysis of
these areas should be prioritized at a regional scale in the Baja Cal- intraspecific differentiation in three endemic bird species from northern
ifornia peninsula. These oases are considered wetlands and human Neotropical dry forest. Mol. Phylogenet. Evol. 70, 362–377.
activity is strongly affecting and changing them, and it can even Arriaga, L., Rodriguez-Estrella, R., Ortega-Rubio, A., 1990. Endemic hummingbird
and madrones of Baja: are they mutually dependent? Southwest Nat. 35, 76–79.
destroy and disappear some of the oases (Rodríguez-Estrella Arriaga, L., Díaz, S., Domínguez, R., León, J.L., 1997. Composición florística y
et al., 2005, 1999). It is clear that if desertification continues, local vegetación. In: Arriaga, L., Rodríguez-Estrella, R. (Eds.), Los Oasis de la Península
extinctions may be inevitable (Gutierrez-Elorza, 1998), but human de Baja California, Mexico. Centro de Investigaciones Biológicas del Noroeste, La
Paz, B.C.S., Mexico, pp. 69–106.
activity can act more than synergistically, because it certainly Arriaga, L., Rodríguez-Estrella, R., 1997. Los Oasis de la Península de Baja California,
change the environment and reduces natural resources that are Mexico. Centro de Investigaciones Biológicas del Noroeste, La Paz, B.C.S.,
already increasingly scarce and vulnerable in desert ecosystems. Mexico.
Avise, J.C., Walker, E., 1998. Pleistocene phylogeographic effects on avian
Oases are small-size, vulnerable habitats. Thus, oases are priority
populations and the speciation process. Proc. R. Soc. Lond. B Biol. Sci. 265,
regions for conservation and from an evolutionary perspective 457–463.
should include not only the conservation of species richness but Avise, J.C., 2000. Phylogeography: The History and Formation of Species. Harvard
of the whole oases as they contain sources of genetic variation that University Press, Cambridge, MA.
Bandelt, H.J., Forster, P., Röhl, A., 1999. Median-joining networks for inferring
support long-term species conservation because of speciation intraspecific phylogenies. Mol. Biol. Evol. 16, 37–48.
process. Berns, C.M., Adams, D.C., 2013. Becoming different but staying alike: patterns of
sexual size and shape dimorphism in hummingbirds. Evol. Biol. 40, 246–260.
Bleiweiss, R., 1998. Tempo a mode of hummingbird evolution. Origin of
5. Conclusions hummingbird faunas. Biol. J. Linn. Soc. 65, 63–97.
Crews, S.C., Hedin, M., 2006. Studies of morphological and molecular phylogenetic
divergence in spiders (Araneae: Homalonychus) from the American southwest,
This study presents the first research of phylogeography of the including divergence along the Baja California Peninsula. Mol. Phylogenet. Evol.
Xantus’ hummingbird, an endemic species of the Baja California 38, 470–487.
Darriba, D., Taboada, G.L., Doallo, R., Posada, D., 2012. JModelTest 2: more models,
Peninsula. Unlike most bird phylogeographical studies in the new heuristics and parallel computing. Nat. Methods 9, 772.
peninsula, the present study, using a saturated sampling strategy De Filippis, V.R., Moore, W.S., 2000. Resolution of phylogenetics relationships a
and three mitochondrial gene sequences, found that Xantus’ hum- recently evolved species as a function of amount DNA sequence: an empirical
study based of woodpeckers (Aves: Picidae). Mol. Phylogenet. Evol. 16, 143–
mingbird showed an ancestral phylogeographic structure with two 160.
different populations along its distribution that began to spread Devitt, T.J., 2006. Phylogeography of the Western Lyresnake (Trimorphodon
during the last 500,000 years. Two different phenomena could biscutatus): testing aridland biogeographical hypotheses across the Nearctic-
Neotropical transition. Mol. Evol. 15, 4387–4407. http://dx.doi.org/10.1111/
explain this phylogeographic pattern: (1) vicariant events such as j.1365-294X.2006.03015.x.
the separation of the peninsula from the mainland, separating H. Drummond, A.J., Rambaut, A., Shapiro, B., Pybus, O.G., 2005. Bayesian coalescent
xantusii from its sister species H. leucotis, and the temporary isola- inference of past population dynamics from molecular sequences. Mol. Biol.
Evol. http://dx.doi.org/10.1093/molbev/msi103.
tion of the southern region at the Isthmus of La Paz, which were
Drummond, A.J., Rambaut, A., 2007. BEAST: Bayesian evolutionary analysis by
supported in this work; and (2) Late Pleistocene climate changes sampling trees. BMC Evol. Biol. 7, 214.
promoting population demographic changes (e.g. expansion Drummond, A.J., Rambaut, A., Suchard, M., 2012. BEAST v1.7.4. Bayesian
event), and together with environmental features and geographic Evolutionary Analysis Sampling Trees <http://beast.bio.ed.ac.uk>.
Dupanloup, I., Schneider, S., Excoffier, L., 2002. A simulated annealing approach to
distribution of oases, leading an incipient genetic differentiation. define the genetic structure of populations. Mol. Ecol. 11, 2571–2581.
This is also the first study that shows two main ancestral refuges Excoffier, L., Lischer, H.E.L., 2010. Arlequin suite ver 3.5: a new series of programs to
for a peninsular endemic species, which highlighted the impor- perform population genetics analyses under Linux and Windows. Mol. Ecol.
Resour. 10, 564–567.
tance of defining areas of evolutionary and conservation relevance Ferrari, L., Conticelli, S., Vaggelli, C., Petrone, C., Manetti, P., 2000. Late Miocene
based on genetic and phylogeographic studies. mafic volcanism and intra-arc tectonics during the early development of the
Trans-Mexican Volcanic Belt. Tectonophysics 318, 161–185.
Fu, Y.X., 1997. Statistical test of neutrality of mutations against populations growth,
Acknowledgements hitchhiking and background selection. Genetics 147, 915–925.
Funk, D.J., Omland, K.E., 2003. Species-level paraphyly and polyphyly: frequency,
causes, and consequences, with insights from animal mitochondrial DNA. Annu.
We thank Franco Cota and José Abelino Cota of CIBNOR and Ser- Rev. Ecol. Evol. Syst. 34, 397–423.
gio Sánchez-González of Universidad Autónoma de Sinaloa for field Galtier, N., Nabholz, B., Glémin, S., Hurst, G.D., 2009. Mitochondrial DNA as a marker
assistance, and Ira Fogel of CIBNOR and Mabilia Urquidi for edito- of molecular diversity: a reappraisal. Mol. Ecol. 18, 4541–4550.
Garrick, R.C., Nason, J.D., Meadows, A., Dyer, R.J., 2009. Not just vicariance:
rial services. We also thank Juan Francisco Ornelas Rodríguez and phylogeography of a Sonoran Desert euphorb indicates a major role of range
Borja Milá for their valuable comments on the use and interpreta- expansion along the Baja peninsula. Mol. Ecol. 18, 1916–1931.
276 C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277

Garrick, R.C., Nason, J.D., Fernández-Manjarrés, Dyer, R.J., 2013. Ecological Chronostratigraphic and Geochronologic Units. US Geological Survey Geologic
coassociations influence species’ responses to past climatic change: an Committee, Washington, D.C.
example from a Sonoran Desert bark beetle. Mol. Ecol. 22, 3345–3361. Polly, D., Speer, B., Patel, S., Chang, R., Jackson, A., Sorenson, P., Wu, Y.Y., Rieboldt, S.,
Gibbard, P., Van Kolfschoten, T., 2004. The pleistocene and holocene epochs. In: Smith, D., 2009. The Pliocene Epoch. International Commission on
Gradstein, F.M., Ogg, J.G., Smith, A.G.A. (Eds.), Geologic Time Scale. Cambridge Stratigraphy’s International Stratigraphic Chart.
University Press (Chapter 22). Qu, Y., Lei, F., 2009. Comparative phylogeography of two endemic birds of the
González, C., Ornelas, J.F., Gutiérrez-Rodríguez, C., 2011. Selection and geographic Tibetan plateu, the white-rumped snow finch (Onychostruthus taczanowskii) and
isolation influence hummingbird speciation: genetic, acoustic and the Hume’s ground tit (Pseudopodoces humilis). Mol. Phylogenet. Evol. 51, 312–
morphological divergence in the wedge-tailed sabrewing (Campylopterus 326.
curvipennis). BMC Evol. Biol. 11, 38. Qu, Y., Zhang, R., Song, G., Zou, F., Lei, F., 2011. Lineage diversification and historical
Grismer, L.L., 2000. Evolutionary biogeography on Mexico’s Baja California demography of montane bird Garrulax elliotii—implications for the Pleistocene
peninsula: a synthesis of molecules and historical geology. Proc. Natl. Acad. evolutionary history of the eastern Himalayas. BMC Evol. Biol. 11, 174.
Sci. USA, doi: 10.1073ypnas.260509697. Rambaut, A., Drummond, A.J., 2009. Tracer v1.6 <http://beast.bio.ed.ac.uk/Tracer>.
Grismer, L.L., McGuire, J.A., 1993. The oases of central Baja California, México. Part I. Rambaut, A., Drummond, A.J., 2012. TreeAnnotator v1.7.4. Institute of Evolutionary
A preliminary account of the relict mesophilic herpetofauna and the status of Biology, University of Edinburgh, Department of Computer Science, University
the oases. Bull. South. Calif. Acad. Sci. 92, 2024. of Auckland <http://beast.bio.ed.ac.uk/>. Available as part of the BEAST package.
Gutierrez-Elorza, M., 1998. Geomorfología Y Cambio Climático En Zonas Áridas. Reding, D.M., Bronikowski, A.M., Johnson, W.E., Clark, W.R., 2012. Pleistocene and
Librería General, Spain. ecological effect on continental-scale genetic differentiation in the bobcat (Lynx
Harpending, H.C., 1994. Signature of ancient population growth in a low-resolution rufus). Mol. Ecol. 21, 3078–3093.
mitochondrial DNA mismatch distribution. Hum. Biol. 66, 591–600. Riddle, B.R., Hafner, D.J., 2006a. A step-wise approach to integrating
Heled, J., Drummond, A.J., 2008. Bayesian inference of population size history from phylogeographic and phylogenetic biogeographic perspectives on the history
multiple loci. BMC Evol. Biol. 8, 289. of the core North American warm deserts biota. J. Arid Environ. 66, 435–461.
Hernández-Baños, B.E., Zamudio-Beltrán, L.E., Eguiarte-Fruns, L.E., Klicka, J., García- Riddle, B.R., Hafner, D.J., Alexander, L.F., 2000a. Phylogeography and systematics of
Moreno, J., 2014. The Basilinna genus (Aves: Trochilidae): an evaluation based the Peromyscus eremicus species group and the historical biogeography of North
on molecular evidence and implications for the genus Hylocharis. Rev. Mex. American warm regional deserts. Mol. Phylogenet. Evol. 17, 145–160.
Biodivers. 85, 797–807. Riddle, B.R., Hafner, D.J., Alexander, L.F., Jaeger, J.R., 2000b. Cryptic vicariance in the
Hewitt, G.M., 1996. Some genetic consequences of ice ages, and their role, in historical assembly of a Baja California Peninsula desert biota. Proc. Nat. Acad.
divergence and speciation. Biol. J. Linn. Soc. 58, 247–276. Sci. USA 97, 14438–14443.
Hewitt, G.M., 2000. The genetic legacy of the Quaternary ice ages. Nature 405, 907– Rodríguez-Estrella, R., 1987. Avifauna. In: Arriaga, L., Ortega, A. (Eds.), La Sierra de la
913. Laguna de Baja California Sur. Centro de Investigaciones Biológicas de Baja
Howell, C.A., Howell, S.N., 2000. Xantus’s Hummingbird (Hylocharis xantusii). The California Sur, México.
Birds of North America, p. 14, online. Rodríguez-Estrella, R., 1997. Factores que condicionan la distribución y abundancia
Hull, J.N., Girman, D.J., 2004. Effects of Holocene climate change on the historical de las aves terrestres en el desierto xerófilo de Baja California Sur, México: El
demography of migrating sharp-shinned hawks (Accipiter striatus velox) in efecto de los cambios en el hábitat por la actividad humana Doctoral Thesis.
North America. Mol. Ecol. http://dx.doi.org/10.1111/j.1365-294X.2004.02366.x. Universidad Autónoma de Madrid-Estación Biológica de Doñana, Spain.
Jensen, J.L., Bohonak, A.J., Kelley, S.T., 2005. Isolation by distance, web service. BMC Rodríguez-Estrella, R., Rubio, L., Pineda, E., 1997. Los oasis como parches atractivos
Genet. 6, 13, v.3.23 <http://ibdws.sdsu.edu/>. para las aves terrestres residentes e invernantes. In: Arriaga, L., Rodríguez-
Jiménez, R.A., Ornelas, J.F., 2016. Historical and current introgression in a Estrella, R. (Eds.), Los Oasis de la Península de Baja California, Mexico. Centro de
Mesoamerican hummingbird species complex: a biogeographic perspective. Investigaciones Biológicas del Noroeste, La Paz, BCS, Mexico, pp. 157–196.
PeerJ. http://dx.doi.org/10.7717/peerj.1556, eCollection. Rodríguez-Estrella, R., Rubio, L., Pineda, E., Blanco, G., 1999. The Belding’s
Kocher, T.D., Thomas, W.K., Meyer, A., Edwards, S.V., Paabo, S., Villablanca, F.X., yellowthroat: current status, habitat preferences, and threats in oases of Baja
Wilson, A.C., 1989. Dynamics of mitochondrial DNA evolution in animals: California, México. Anim. Conserv. 2, 77–84.
amplification and sequencing with conserved primers. Proc. Natl. Acad. Sci. USA Rodríguez-Estrella, R., Blázquez, M.C., Lobato, J.M., 2005. Avian communities of
86, 6196–6200. arroyos and desert oases in Baja California Sur: implications for conservation.
Lara, C., Feria- Arroyo, T.P., Dale, J., Muñoz, J., del Coro-Arizmendi, M., Ornelas, J.F., In: Cartron, J.L.E., Ceballos, G., Felger, R.S. (Eds.), Biodiversity and Conservation
Ortiz-Pulido, R., Rodríguez-Flores, C.I., Díaz-Valenzuela, R., Martínez-García, V., in Northern Mexico. Oxford University Press, UK, pp. 334–353.
Díaz-Palacios, A., Partida, R., Enríquez, P.L., Rangel-Salazar, J.L., Schondube, J., Rodríguez-Gómez, F., Ornelas, J.F., 2015. At the passing gate: past introgression in
2012. Potential effects of the climate change in the distribution of the process of species formation between Amazilia violiceps and A. viridifrons
hummingbirds: a study case with hummingbirds from the genus Amazilia and hummingbirds along the Mexican Transition Zone. J. Biogeogr. 42, 1305–1318.
Cynanthus. Ornitol. Neotrop. 23, 57–70. http://dx.doi.org/10.1111/jbi.12506.
Leaché, A.D., Mulcahy, D.G., 2007. Phylogeny, divergence times and species limits of Rogers, A.R., Harpending, H., 1992. Population growth makes waves in the
spiny lizards (Sceloporus magister species group) in western North American distribution of pairwise genetic differences. Mol. Biol. Evol. 9, 552–569.
deserts and Baja California. Mol. Ecol. 16, 5216–5233. Rojas-Soto, O.R., Alcántara-Ayala, O., Navarro, A.G., 2003. Regionalization of the
Leaché, A.D., Koo, M.S., Spencer, C.L., Papenfuss, T.J., Fisher, R.N., McGuire, J.A., 2009. avifauna of the Baja California Peninsula, Mexico: a parsimony analysis of
Quantifying ecological, morphological, and genetic variation to delimit species endemicity and distributional modelling approach. J. Biogeogr. 30, 449–461.
in the coast horned lizard species complex (Phrynosoma). Proc. Natl. Acad. Sci. Ronquist, F., Teslenko, M., Van der Mark, P., Ayres, D.L., Darling, A., Höhna, S., Larget,
106, 12418–12423. B., Liu, L., Suchard, M.A., Huelsenbeck, J.P., 2012. Mrbayes 3.2: efficient Bayesian
Lemmon, E.M., Lemmon, A.R., Collins, J.T., Lee-Yaw, J.A., Cannatella, D.C., 2007. phylogenetic inference and model choice across a large model space. Syst. Biol.
Phylogeny-based delimitation of species boundaries and contact zones in the 61, 539–542.
trilling chorus frogs (Pseudacris). Mol. Phylogenet. Evol. 44, 1068–1082. Scoble, J., Lowe, A.J., 2010. A case for incorporating phylogeography and landscape
Librado, P., Rozas, J., 2009. DnaSP v5: a software of comprehensive analysis of DNA genetics into species distribution modeling approaches to improve climate
polymorphism data. Bioinformatics 25, 1451–1452. adaptation and conservation planning. Divers. Distrib. 16, 343–353.
Licona-Vera, Y., Ornelas, J.F., 2014. Genetic, ecological and morphological Sly, N.D., Townsend, A.K., Rimmer, C.C., Townsend, J.M., Latta, S.C., Lovette, I.J., 2011.
divergence between populations of the endangered Mexican sheartail Ancient islands and modern invasions: disparate phylogeographic histories
hummingbird (Doricha eliza). PLoS ONE 9, e101870. http://dx.doi.org/10.1371/ among Hispaniola’s endemic birds. Mol. Ecol. 20, 5012–5024.
journal.pone.0101870. Smith, B.T., Escalante, P., Hernández-Baños, B.E., Navarro-Sigüenza, A.G., Rohwer, S.,
Mapelli, F.J., Mora, M.S., Mirol, P.M., Kittlein, M.J., 2010. Effects of Quaternary Klicka, J., 2011. The role of historical and contemporary processes on
climatic changes on the phylogeography and historical demography of the phylogeographic structure and genetic diversity in the northern cardinal,
subterranean rodent Ctenomys porteousi. J. Zool. http://dx.doi.org/10.1111/ Cardinalis cardinalis. BMC Evol. Biol. 11, 136.
j.1469-7998.2011.00849.x. Sorenson, M.D., Ast, J.C., Dimcheff, D.E., Yuri, T., Mindell, D.P., 1999. Primers for a
Maya, Y., Coria, R., Domínguez, R., 1997. Caracterización de los Oasis. In: Arriaga, L., PCR-based approach to mitochondrial genome sequencing in birds and other
Rodríguez-Estrella, R. (Eds.), Los Oasis de la Península de Baja California, vertebrates. Mol. Phylogenet. Evol. 12, 105–114.
Mexico. Centro de Investigaciones Biológicas del Noroeste, La Paz, B.C.S., Tajima, F., 1989. Statistical method for testing the neutral mutation hypothesis by
Mexico, pp. 5–26. DNA polymorphism. Genetics 123, 585–595.
McGuire, J.A., Witt, C.C., Altshuler, D.L., Remsen Jr., J.V., 2007. Phylogenetic Tamura, K., Nei, M., 1993. Estimation of the number of nucleotide substitutions in
systematics and biogeography of hummingbirds: Bayesian and maximum the control region of mitochondrial DNA in humans and chimpanzees. Mol. Biol.
likelihood analyses of partitioned data and selection of an appropriate Evol. 10, 512–526.
partitioning strategy. Syst. Biol. 56, 837–856. Tamura, K., Stecher, G., Peterson, D., Filipski, A., Kumar, S., 2013. MEGA6: molecular
McGuire, J.A., Witt, C.C., Remsen Jr., J.V., Corl, A., Rabosky, D.L., Altshuler, D.L., evolutionary genetic analysis version 6.0. Mol. Biol. Evol. 30, 2725–2729.
Dudley, 2014. Molecular phylogenetics and the diversification of Temeles, E.J., Miller, J.S., Rifkin, J.L., 2010. Evolution of sexual dimorphism in bill size
hummingbirds. Curr. Biol. http://dx.doi.org/10.1016/j.cub.2014.03.016. and shape of hermit hummingbirds (Phaethornithinae): a role for ecological
Munguia-Vega, A., 2011. Habitat Fragmentation in Small Vertebrates from the causation. Phil. Trans. R. Soc. B 365, 1053–1063.
Sonoran Desert in Baja California Doctoral Thesis. University of Arizona, Tucson, Valero, A., Schipper, J., Allnutt, T., 2014. Southern North America: Baja California
AZ. Peninsula in Western Mexico http://www.worldwildlife.org/ecoregions/na1306.
Orndorff, R.C., Stamm, N., Craigg, S., D’Erchia, T., Edwards, L., Fullerton, D., Murchey, Weir, J.T., Schluter, D., 2008. Calibrating the avian molecular clock. Mol. Ecol. 17,
B., Ruppert, L., Soller, D., Tew, B.N., 2007. Divisions of Geologic Time. Major 2321–2328.
C. González-Rubio et al. / Molecular Phylogenetics and Evolution 102 (2016) 265–277 277

Wiens, J.J., 2015. Explaining large-scale patterns of vertebrate diversity. Biol. Lett. Zink, R.M., Blackwell-Rago, R.C., Ronquist, F., 2000. The shifting roles of dispersal
11, 20150506. http://dx.doi.org/10.1098/rsbl.2015.0506. and vicariance in biogeography. Proc. R. Soc. Lond. B Biol. Sci. 267, 497–503.
Wright, S., 1943. Isolation by distance. Genetics 28, 114–138. Zink, R.M., Kessen, A.E., Line, T.V., Blackwell-Rago, R.C., 2001. Comparative
Zamudio-Beltrán, L.E., 2011. Filogenia y variación dentro del complejo Hylocharis phylogeography of some aridland bird species. Condor 103, 1–10.
leucotis Vieillot, 1818 (Aves: Trochilidae) utilizando secuencias de DNA Master’s Zink, R.M., Groth, J.G., Vázquez-Miranda, H., Barrowclough, G.F., 2013.
Thesis. Facultad de Ciencias, Universidad Nacional Autónoma de México, Phylogeography of the California gnatcatcher (Polioptila californica) using
Mexico City. multilocus DNA sequences and ecological niche modeling: implications for
Zink, R.M., 2002. Methods in comparative phylogeography, and their application to conservations. Auk 130, 449–458.
studying evolution in the North American arid lands. Integr. Comp. Biol. 42, Zwickl, D.J., 2011. GARLI 2.0 https://www.nescent.org/wg_garli/main_page.
953–959.

You might also like