ArticlePDF Available

The Influence of Pleistocene Refugia on the Evolutionary History of the Japanese Hare, Lepus brachyurus

Authors:
  • Oyakama University of Science

Abstract and Figures

We performed a phylogeographic analysis of the Japanese hare, Lepus brachyurus, using the mitochondrial cytochrome b gene (1140 bp). In total, 119 haplotypes were recovered from 197 samples isolated from 82 localities on three main islands of the Japanese archipelago: Honshu, Sikoku, Kyushu, Sado Island and the Oki Islands. Results showed two distinct clades at a genetic distance of 3.5%, equivalent to an estimated 1.2 million years. The two clades, encompassing seven subclades, showed an apparent geographic affinity to Kyushu, Shikoku and the nearby area of Honshu (southern group) by one clade, whereas the other clade covered the remaining area of Honshu (northern group). The landscape shape interpolation analysis exhibited a higher genetic diversity in the southern parts of central Honshu (northern group) and Shikoku and Kyushu regions (southern group), suggesting the existence of multiple geographical origins of population expansion in each clade. The Bayesian skyline plot analysis showed that lineage diversifications occurred about 0.35, 0.20 and 0.05 million years ago (Mya), which coincide closely with the glacial-interglacial cycles during the Pleistocene. Therefore, we suggest that the Japanese hare population once inhabited northern and southern refugia, and subsequently developed several populations through local demographic fluctuations. The present day demarcation in the northern and southern geographic groups is considered to be a temporal remnant of Pleistocene population dynamics and the geographic boundary between them could move or fade away in time.
Content may be subject to copyright.
Lack of association between winter coat colour and
genetic population structure in the Japanese hare,
Lepus brachyurus (Lagomorpha: Leporidae)
MITSUO NUNOME1*, GOHTA KINOSHITA2, MORIHIKO TOMOZAWA3, HARUMI TORII4,
RIKYU MATSUKI5, FUMIO YAMADA6, YOICHI MATSUDA1and HITOSHI SUZUKI2
1Laboratory of Animal Genetics, Department of Applied Molecular Biosciences, Graduate School of
Bioagricultural Sciences, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8601, Japan
2Laboratory of Ecology and Genetics, Faculty of Environmental Earth Science, Hokkaido University,
Kita-ku, Sapporo 060-0810, Japan
3Department of Biology, Keio University, Yokohama 223-8521, Japan
4Center for Natural Environment Education, Nara University of Education, Takabatake-cho, Nara
630-8528, Japan
5Environmental Science Research Laboratory, Central Research Institute of Electric Power Industry,
1646 Abiko, Chiba 270-1194, Japan
6Forestry and Forest Products Research Institute, PO Box 16, Tsukuba Norin, Ibaraki 305-8687,
Japan
Received 3 October 2013; revised 11 December 2013; accepted for publication 11 December 2013
Seasonal changes in fur colour in some mammalian species have long attracted the attention of biologists,
especially in species showing population variation in these seasonal changes. Genetic differences among popula-
tions that show differences in seasonal changes in coat colour have been poorly studied. Because the Japanese hare
(Lepus brachyurus) has two allopatric morphotypes that show remarkably different coat colours in winter, we
examined the population genetic structure of the species using partial sequences of the SRY gene and six autosomal
genes: three coat colour-related genes (ASIP,TYR, and MC1R) and three putatively neutral genes (TSHB,APOB,
and SPTBN1). The phylogenetic tree of SRY sequences exhibited two distinct lineages that diverged approsimately
1 Mya. Although the two lineages exhibited a clear allopatric distribution, it was not consistent with the
distribution of morphotypes. In addition, six nuclear gene sequences failed to reveal genetic differences between
morphotypes. Population network trees for 11 expedient populations divided the populations into four groups.
Genetic structure analysis revealed an admixture of four genetic clusters in L. brachyurus, two of which showed
large genetic differences. Our results suggest ancient vicariance in L. brachyurus, and we detected no genetic
differences between the two morphotypes. © 2014 The Linnean Society of London, Biological Journal of the
Linnean Society, 2014, 111, 761–776.
ADDITIONAL KEYWORDS: adaptation – natural selection – phylogeography – seasonal change.
INTRODUCTION
The coat colours of some mammalian and avian
species that live in middle to high latitudinal areas,
such as grouses, foxes, martens, weasels, hares, and
hamsters, change seasonally, from dark colours in
summer to light colours in winter. Such seasonal
changes in coat colour have attracted the interest of
numerous researchers for many years (Grange, 1932;
Hewson, 1958; Rust, 1965; Flux, 1970; Watson, 1973;
Walsberg, 1991; Russell & Tumlison, 1996; Stoner,
Bininda-Emonds & Caro, 2003; Scherbarth &
Steinlechner, 2010). Morphological studies have been
*Corresponding author.
E-mail: mtnunome@agr.nagoya-u.ac.jp
bs_bs_banner
Biological Journal of the Linnean Society, 2014, 111, 761–776. With 4 figures
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776 761
performed in hares aiming to understand seasonal
variation in fur colour and to identify environmental
cues that induce the change (Hewson, 1958; Otsu,
1967; Flux, 1970; Kuderling et al., 1984). These
studies have indicated that day length is a definite
signal for changes in coat colour, and that tempera-
ture is another important cue for the timing of these
changes (Mills et al., 2013). Hare species also change
reproductive activities in response to day length. In
addition, some hare species show intraspecific varia-
tion in the seasonality of coat colour. In such species,
animals in northern areas show seasonal coat colour
changes, whereas those in southern areas retain dark
pelages year-round.
Intraspecific variation in winter coat colour is con-
sidered to originate from genetic differences among
populations. This hypothesis is supported by various
observations as described below. However, the genes
that regulate seasonal changes in coat colour have
not yet been identified in hares or other animals.
Although variation in the melanocortin 1 receptor
(MC1R) gene is known to be associated with variation
in winter coat colour in the Arctic fox (Alopex lagopus)
(Vage et al., 2005), the gene is not involved in the
regulation of seasonal changes in coat colour in the
Japanese marten (Martes melampus) (Hosoda et al.,
2005) or the willow grouse (Skoglund & Hoglund,
2010). Moreover, few studies have examined genetic
differences between populations that show differences
in seasonal moult (Meinke, Kapel & Arctander, 2001;
Sato, Yasuda & Hosoda, 2009). In these studies, no
remarkable genetic differences were found between
the two morphotypes. In Lepus species, which are
well known to show seasonal moult, the genetic
basis of morphotypic differences has not been well
studied, although numerous morphological, physi-
ological, and phylogeographical studies have been
conducted (Grange, 1932; Hewson, 1958; Watson,
1963; Flux, 1970; Kuderling et al., 1984; Iason &
Ebling, 1989; Koutsogiannouli et al., 2012). Thus,
uncovering the genetic basis of the two morphotypes
in hare species is crucial not only for identifying the
gene(s) involved in seasonal moult, but also for under-
standing how differences in winter coat colour are
related to population genetic structure.
The Japanese hare Lepus brachyurus Temminck,
1845 is an appropriate model in which to examine the
genetic basis of winter colour morphotypes because of
its restricted geographical distribution and genetic
status. Lepus brachyurus is endemic to the Japanese
archipelago and is distributed throughout three of
the main islands (Honshu, Shikoku, and Kyushu),
as well as in peripheral islands, including Sado
Island and the Oki Islands (Fig. 1). The species has a
much smaller range than other Lepus species that are
found on the continent. In addition, although several
hare species are known to have experienced genetic
introgression from other species (Alves et al., 2008;
Liu et al., 2011; Kinoshita et al., 2012), mitochondrial
phylogenetic studies have confirmed the genetic
purity of L. brachyurus (Yamada, Takaki & Suzuki,
2002; Wu et al., 2005; Nunome et al., 2010). Thus,
compared to other hare species on the continent,
the examination of genetic population structure
is relatively simple in almost all populations of
L. brachyurus.
Lepus brachyurus is divided principally into north-
ern and southern groups based on winter coat colour
(Fig. 1). The coat colours of hares in northern areas
change from brown in summer to white in winter,
whereas hares in southern areas have brown pelages
year-round (Imaizumi, 1960; Hirata, 1999). This
intraspecific difference in winter pelage has led to the
consideration of the northern and southern groups as
subspecies; white winter pelage animals are desig-
nated Lepus brachyurus angustidens Hollister, 1912,
and pigmented winter pelage animals are classified
as Lepus brachyurus brachyurus Temminck, 1845,
excluding two subspecies on Sado Island and the Oki
Islands. Hereafter, for descriptive purposes, hares
with white and pigmented winter coats on the main
islands are referred to as ‘Lba’ and ‘Lbb’, respectively.
When Lba is bred in areas where Lbb is dominant,
Lba changes its coat colour from brown in summer to
white in winter (Hirata, 1999), suggesting that the
seasonal change in coat colour of Lba is not triggered
by environmental cues but, instead, is determined by
genetic factor(s). The distribution ranges of Lba and
Lbb do not display simple north–south parapatry but
are strongly related to annual snowfall in the archi-
pelago. On Honshu Island, the amount of annual
snowfall differs distinctly between the heavy-snow
region on the Japan Sea side and the low-snow region
on the Pacific side as a result of the presence of high
mountain chains that run north–south through the
middle of the island. Winter coat colour may provide
cryptic coloration in snowy environments. That is,
clear geographical differences in annual snowfall
within the archipelago strongly affect the ranges of
the two morphotypes, resulting into genetic diver-
gence between them.
A previous phylogeographical study of mitochondrial
DNA variation in L. brachyurus found two distinct
lineages. However, the geographical distributions of
the lineages were not consistent with the distributions
of the morphotypes, and genetic differences between
the morphotypes remain to be examined (Fig. 2)
(Nunome et al., 2010). In the present study, we exam-
ined genetic variation in L. brachyurus using multiple
nuclear DNA sequences to clarify genetic differences
between the two morphotypes. First, a phylogene-
tic tree of L. brachyurus was constructed using the
762 M. NUNOME ET AL.
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
Y-linked, sex-determining gene SRY to confirm the two
distinct lineages of the species that were exhibited in
cytochrome b gene (CYT B) sequences in a previous
study (Nunome et al., 2010). Second, to determine
whether any of the three coat colour genes is the basis
of the two morphotypes of L. brachyurus, we assessed
genetic variation in six nuclear DNA loci among
11 expedient populations. Three of the six markers
were thyrotropin beta subunit (TSHB), beta spectrin
(SPTBN1), and apolipoprotein B 100 (APOB), which
have been used widely as neutral markers in
phylogenetic analyses of mammals, including Lepus
species (TSHB and SPTBN1: Matthee et al., 2004;
Hoofer et al., 2008; APOB: Amrine-Madsen et al.,
2003). The other three markers were agouti signalling
protein (ASIP), Tyrosinase (TYR) and MC1R, which
are known to be responsible for light and dark coat-
colour variations in mammalian species, including
lagomorphs (Chen, Duhl & Barsh, 1996; Aigner et al.,
2000; Miltenberger et al., 2002; Nachman, Hoekstra &
D’Agostino, 2003; Kambe et al., 2011). Third, we con-
ducted population genetic structure analyses with
the six autosomal loci to evaluate genetic differences
between the two morphotypes.
MATERIAL AND METHODS
SAMPLING AND DNA EXTRACTION
Total genomic DNA was extracted from skin, liver
(sampled from road-killed or hunted animals) and
faeces using a traditional phenol/chloroform protocol
(Sambrook & Russell, 2001). Tissue samples of
L. brachyurus were obtained from 50 localities
across the distribution area (Fig. 1, Table 1). Partial
sequences of genes were amplified for one Y-linked
gene (SRY) and six autosomal genes (ASIP,MC1R,
TYR,APOB,SPTBN1, and TSHB) using the polymer-
ase chain reaction (PCR) method. Amplifications
were performed in a total volume of 20 μL containing
approximately 10 ng of genomic DNA, 10 pmol of each
primer and 10 μL of AmpliTaq Gold®360 Master Mix
(Life Technologies). Cycling conditions for PCR were:
initial denaturation at 95 °C for 10 min, followed by
35 cycles at 95 °C for 30 s, 52 °C–62 °C for 30 s, and
70 °C for 30 s. The primers and exact annealing tem-
perature for each locus are shown in the Supporting
information (Table S1). Double-stranded PCR prod-
ucts were purified using the 20% polyethylene glycol/
2.5 M NaCl precipitation method. The PCR products
10°N
20°
30°
40°
50°
90° 100° 110° 120° 130°80°E 140°
17
18 16
14
13
12
25
24
22
28
29
31
32
33
35
38
36 37
41
43
45 48
46
39
40
10
4
67
9
8
20
21
23
19
15
11
26
27
30
34
42
44
49
50
1
5
23
47
Honshu
Tohoku
Kanto
Chubu
Kinki
Shikoku
Kyushu
Chugoku
Sado isl.
Oki isls.
pop1
pop4
pop9
pop8
pop10
pop7
pop2
pop6 pop5
pop3
pop11
Figure 1. Map of sampling locations. The darkly shaded area indicates the distribution area of animals that have white
pelages in winter. Samples were divided into 11 populations according to their geographical locations and phylogenetic
trees of CYT B and SRY.
POPULATION GENETIC STRUCTURE IN JAPANESE HARE 763
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
were sequenced from both directions with a BigDye
Terminator cycle sequencing kit, version 3.1 (Life
Technologies) and analyzed using an ABI 3100
automated sequencer (Applied Biosystems). DNA
sequences were aligned visually using PROSEQ,
version 2.9.1 (Filatov, 2001).
PHYLOGENETIC ANALYSIS AND DIVERGENCE TIME
ESTIMATION FOR SRY
A Neighbour-joining (NJ) tree (Saitou & Nei, 1987) was
constructed using PAUP 4.0b10 (Swofford, 2002). As a
distance model for the NJ tree, the Hasegawa–
Kishino–Yano (HKY) distance model was chosen based
on a hierarchical likelihood ratio test implemented
with MODELTEST, version 3.7 (Posada & Crandall,
1998). Published SRY sequences (DDBJ/EMBL/
GenBank) for the European hare (L. europaeus;
EF437189, EF437190) were used as an outgroup.
One thousand bootstrap replicates were performed
to assess the robustness of each node. A maximum-
likelihood tree was reconstructed using PHYML,
version 3.0 (Guindon & Gascuel, 2003; Guindon et al.,
2005). Node robustness in the tree was evaluated using
100 bootstrap replicates. Divergence times between
Japanese hare lineages were inferred using BEAST,
version 1.4.8 (Drummond et al., 2005). SRY sequences
for rabbit (Oryctolagus cuniculus; AY785433) were
included in the dataset to set a calibration point
for estimating divergence times. The divergences
between O. cuniculus and the Lepus group, and
between L. europaeus and L. brachyurus, were set
at 11 Mya and 3.5 Mya, respectively, based on esti-
mates from previous molecular phylogenetic studies
(Matthee et al., 2004; Wu et al., 2005). The HKY with
the SRD06 model was used as a substitution model.
Analyses were run for 10 million generations, with
sampling conducted every 1000 generations following
one million burn-in generations. Convergence was
assessed using TRACER, version 1.5 (Rambaut &
Drummond, 2007).
SEQUENCE ANALYSIS FOR SIX AUTOSOMAL LOCI
The minimum numbers of recombinations (Rms;
Hudson & Kaplan, 1985) in the sequences of six
autosomal genes were examined using DNASP,
version 5.0.0 (Librado & Rozas, 2009), and the longest
18
24
32
33
38
36 37
41
39
67
8
15
30
46
47
35
17
16
12
22
23
1
3
0.001
8_NIG218
47_OHIT88
46_FKOK85
23_TYM129
EF437189LU
6_GNM225
17_IBRK82
41_KCH136
38_SHMN36
3_AKT238
18_TCG101
18_TCG102
32_NARA57
22_TYM128
36_HRSM25
36_HRSM27
7_GNM216
3_AKT242
37_HRSM32
24_SZOK04
12_IBRK71
15_IBRK83
16_IBRK81
35_AWJ_18
30_MIE_70
1_AKT251
EF437190LU
38_SHMN39
6_GNM214
39_OKNS43
33_NARA84
33_NARA59
95/95
63/62
95/97
62/64
lineage N
lineage S
outgroup
Figure 2. Phylogenetic tree of partial DNA sequences of SRY. Codes at the tips of the tree represent the numbers of
sampling locations, followed by the sample codes. Values on branches indicate bootstrap values for the Neighbour-joining
(left) and maximum-likelihood (right) methods. The geographical border between the two lineages in the tree is indicated
by a dotted line on the map of sampling localities. The grey dotted line on the map indicates the geographical border
between two clades of the mitochondrial CYT B gene (Nunome et al., 2010).
764 M. NUNOME ET AL.
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
Table 1. Details of samples used in the present study
Population
number
Sample
name
Locality
number District Prefecture City (Gun) Latitude Longitude
1 LbAKT251 1 Tohoku Akita Kitaakita 140.37 40.23
LbAKT235 2 Tohoku Akita Nikaho 139.91 39.20
LbAKT236 2 Tohoku Akita Nikaho 139.91 39.20
LbAKT237 3 Tohoku Akita Yuzawa 140.49 39.16
LbAKT238 3 Tohoku Akita Yuzawa 140.49 39.16
LbAKT239 3 Tohoku Akita Yuzawa 140.49 39.16
LbAKT240 3 Tohoku Akita Yuzawa 140.49 39.16
LbAKT241 3 Tohoku Akita Yuzawa 140.49 39.16
LbAKT242 3 Tohoku Akita Yuzawa 140.49 39.16
LbAKT243 3 Tohoku Akita Yuzawa 140.49 39.16
LbIWT205 4 Tohoku Iwate Shimohei 141.87 39.69
LbIWT206 5 Tohoku Iwate Morioka 141.19 39.67
2 LbGNM210 6 Kanto Gunma Azuma 138.50 36.50
LbGNM211 6 Kanto Gunma Azuma 138.50 36.50
LbGNM212 6 Kanto Gunma Azuma 138.50 36.50
LbGNM214 6 Kanto Gunma Azuma 138.50 36.50
LbGNM225 6 Kanto Gunma Azuma 138.50 36.50
LbGNM215 7 Kanto Gunma Tone 139.05 36.73
LbGNM216 7 Kanto Gunma Tone 139.05 36.73
LbGNM217 7 Kanto Gunma Tone 139.05 36.73
LbNIG218 8 Chubu Niigata Tokamachi 138.76 37.13
LbNIG219 8 Chubu Niigata MinamiUonuma 138.82 36.95
LbNIG220 9 Chubu Niigata MinamiUonuma 138.82 36.95
LbSAD189 10 Sado isl. Niigata Sado 138.37 38.02
LbSAD190 10 Sado isl. Niigata Sado 138.37 38.02
LbSAD191 10 Sado isl. Niigata Sado 138.37 38.02
3 LbIBRK55 20 Chubu Nagano KitaSaku 138.55 36.33
LbIBRK77 20 Chubu Nagano KitaSaku 138.55 36.33
LbIBRK71 20 Chubu Nagano KitaSaku 138.55 36.33
LbIBRK73 20 Chubu Nagano KitaSaku 138.55 36.33
LbIBRK75 21 Chubu Nagano Shimotakai 138.43 36.81
LbIBRK76 21 Chubu Nagano Shimotakai 138.43 36.81
LbIBRK78 22 Chubu Toyama Toyama 137.21 36.70
LbIBRK83 23 Chubu Toyama Oyabe 137.00 36.68
4 LbIBRK81 11 Kanto Ibaraki Tsukuba 140.07 36.04
LbIBRK82 11 Kanto Ibaraki Tsukuba 140.07 36.04
LbTCG100 12 Kanto Ibaraki Bando 139.89 36.04
LbTCG101 13 Kanto Ibaraki 140.28 36.25
LbTCG102 14 Kanto Ibaraki Kasama 140.24 36.38
LbTCG103 14 Kanto Ibaraki Kasama 140.24 36.38
LbTCG104 15 Kanto Ibaraki Kasumigaura 140.32 36.08
LbTCG106 15 Kanto Ibaraki Kasumigaura 140.32 36.08
LbTCG107 16 Kanto Ibaraki Mito 140.45 36.37
LbTCG209 17 Kanto Ibaraki Takahagi 140.70 36.72
LbNGN221 18 Kanto Tochigi Nikko 139.60 36.74
LbNGN222 18 Kanto Tochigi Nikko 139.60 36.74
LbNGN223 18 Kanto Tochigi Nikko 139.60 36.74
LbNGN224 18 Kanto Tochigi Nikko 139.60 36.74
LbNGN226 18 Kanto Tochigi Nikko 139.60 36.74
LbNGN227 18 Kanto Tochigi Nikko 139.60 36.74
LbTYM128 18 Kanto Tochigi Nikko 139.60 36.74
LbTYM129 19 Kanto Tochigi Nasushiobara 140.05 36.96
POPULATION GENETIC STRUCTURE IN JAPANESE HARE 765
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
Table 1. Continued
Population
number
Sample
name
Locality
number District Prefecture City (Gun) Latitude Longitude
5 LbSZOK02 24 Chubu Shizuoka Fujinomiya 138.62 35.22
LbSZOK04 24 Chubu Shizuoka Fujinomiya 138.62 35.22
LbSZOK07 25 Chubu Shizuoka Susono 138.91 35.17
LbSZOK09 25 Chubu Shizuoka Susono 138.91 35.17
LbYMN130 26 Chubu Yamanashi Fujiyoshida 138.81 35.49
LbYMN131 27 Chubu Yamanashi Minamitsuru 138.80 35.48
LbYMN134 27 Chubu Yamanashi Minamitsuru 138.80 35.48
6 LbMIE_68 28 Kinki Mie Inabe 136.56 35.12
LbMIE_69 29 Kinki Mie Tsu 136.51 34.72
LbMIE_70 30 Kinki Mie 136.51 34.73
LbNARA12 31 Kinki Nara Odaigahara 135.88 34.39
LbNARA57 32 Kinki Nara Katsuragi 135.71 34.51
LbNARA59 33 Kinki Nara Nara 135.81 34.69
LbNARA84 33 Kinki Nara Nara 135.81 34.69
LbSHIG97 34 Kinki Shiga Otsu 135.85 35.02
7 LbAWJ_17 35 Chugoku Hyogo Awaji 134.92 34.44
LbAWJ_18 35 Chugoku Hyogo Awaji 134.92 34.44
LbAWJ_19 35 Chugoku Hyogo Awaji 134.92 34.44
LbAWJ_20 35 Chugoku Hyogo Awaji 134.92 34.44
LbAWJ_21 35 Chugoku Hyogo Awaji 134.92 34.44
8 LbHRSM25 36 Chugoku Hiroshima Mihara 133.08 34.40
LbHRSM26 36 Chugoku Hiroshima Mihara 133.08 34.40
LbHRSM27 36 Chugoku Hiroshima Mihara 133.08 34.40
LbHRSM28 36 Chugoku Hiroshima Mihara 133.08 34.40
LbHRSM32 37 Chugoku Hiroshima Onomichi 133.20 34.41
LbHRSM34 37 Chugoku Hiroshima Onomichi 133.20 34.41
LbSHMN36 38 Chugoku Shimane Nita 133.05 35.18
LbSHMN37 38 Chugoku Shimane Nita 133.05 35.18
LbSHMN39 38 Chugoku Shimane Nita 133.05 35.18
9 LbOKI_42 39 Oki isl. Shimane Nishinoshima 132.98 36.10
LbOKI_43 39 Oki isl. Shimane Nishinoshima 132.98 36.10
LbOKI_44 39 Oki isl. Shimane Nishinoshima 132.98 36.10
LbOKI_46 40 Oki isl. Shimane Dogo 133.28 36.25
LbOKI_47 40 Oki isl. Shimane Dogo 133.28 36.25
LbOKI_49 40 Oki isl. Shimane Dogo 133.28 36.25
LbOKI_53 40 Oki isl. Shimane Dogo 133.28 36.25
LbOKI_54 40 Oki isl. Shimane Dogo 133.28 36.25
10 LbKCH136 41 Shikoku Kochi Nagaoka 133.65 33.77
LbKCH138 42 Shikoku Kochi Tosa (gun) 133.53 33.74
LbKCH140 42 Shikoku Kochi Tosa (gun) 133.53 33.74
LbKCH229 43 Shikoku Kochi Kami 133.69 33.60
LbKCH230 44 Shikoku Kochi Tosa (city) 133.43 33.50
11 LbFKO163 45 Kyushu Fukuoka Kama 130.77 33.56
LbFKOK85 46 Kyushu Fukuoka KitaKyushu 130.88 33.88
LbOITA86 47 Kyushu Oita Beppu 131.49 33.28
LbOITA88 47 Kyushu Oita Beppu 131.49 33.28
LbOITA92 48 Kyushu Oita Usa 131.33 33.53
LbOITA87 49 Kyushu Oita Yufu 131.43 33.18
LbOITA91 49 Kyushu Oita Yufu 131.43 33.18
LbOITA93 49 Kyushu Oita Yufu 131.43 33.18
LbKGS233 50 Kyushu Kagoshima Satsumasendai 130.30 31.81
LbKGS234 50 Kyushu Kagoshima Satsumasendai 130.30 31.81
Sampling locations are represented by the city or area where the samples were collected. The longitude and latitude of
each sampling location were determined using Google Earth 6.02; (Google, Inc.).
766 M. NUNOME ET AL.
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
portions without recombination signals were used in
the present study. We subdivided the MC1R sequence
into two parts (MC1R-a and MC1R-b) based on the
results of the recombination test. Tajima’s neutrality
tests (Tajima’s D; Tajima, 1989) and linkage disequi-
librium tests were performed with 10 000 permuta-
tions using ARLEQUIN, version 3.5 (Excoffier &
Lischer, 2010). We subdivided the samples into 11
populations according to their geographical locations
and phylogenetic data of previous CYT B sequences
and the SRY sequences from this study (populations
1–11; Fig. 1). First, samples were partitioned accord-
ing to the three main islands of the Japanese archi-
pelago (Honshu, Shikoku, and Kyushu). Samples from
the Oki Islands were treated as one population. Then,
samples from Honshu were further subdivided based
on the borders of the two mitochondrial lineages
(population 8 and the others), the two SRY lineages
(such as populations 2 and 3), the two morphotypes
(such as populations 3 and 5) and geographical dis-
tances (such as populations 1 and 2). Geographical
variations in the six genes were surveyed by examin-
ing haplotype frequencies in 11 populations. Isolation-
by-distance (IBD) was evaluated for each gene dataset
using a Mantel test with 1000 replicates using
ALLELE IN SPACE (Miller, 2005). Genetic diversity
[expected heterozygosity (HE) and pairwise nucleotide
differences (pi)] within populations for each locus were
examined using ARLEQUIN, verison 3.5.
POPULATION NETWORK CONSTRUCTION AND
POPULATION GENETIC STRUCTURE ANALYSIS
To infer genetic differences between the two morpho-
types, exact tests (Raymond & Rousset, 1995) were
performed using ARLEQUIN, version 3.5, for two
representative groups of populations (populations 1–3
and 4–11), based on the distributions of Lba and Lbb.
In addition, according to two CYT B lineages, an
exact test was performed for two other population
groupings (populations 1–6 and 7–11).
Genetic differences among populations were also
examined through two independent analyses. First,
we constructed population network trees to determine
the genetic relationships among the 11 populations,
with a special focus on significant differences in
the pairwise fixation index (FST) among populations.
Second, population genetic structure was inferred
using STRUCTURE, version 2.3 (Pritchard, Stephens
& Donnelly, 2000) to determine the genetic back-
grounds of the 11 populations. Genetic relationships
among populations have commonly been constructed
as unrooted trees based on pairwise FST (Wright,
1951) or Nei’s genetic distance (Nei’s DA; Nei & Li,
1979). However, these trees are rectilinear and cannot
present significant genetic differences among popula-
tions. Thus, to survey population genetic relation-
ships and identify significant genetic differences, we
used significant FST distances in network reconstruc-
tion in addition to performing conventional network
reconstruction using pairwise FST distances. First,
pairwise FST differences between populations and
their corresponding Pvalues were calculated for the
six nuclear genes using ARLEQUIN, version 3.5, with
1000 permutation replicates. Then, a conventional
FST network based on the NJ method was recon-
structed using TREEFIT (Kalinowski, 2009). Second,
we assigned values of ‘1’ or ‘0’ to Pvalues in the
resulting data matrix according to their significance
(P<0.05) or nonsignificance (P0.05), respectively.
We used ‘0’ (meaning nonsignificant difference) for
within-population values, which were diagonal blank
elements in the Pvalue matrix. Finally, a binary
alignment of 0 and 1 for each population was gener-
ated. Then, for the binary data, reduced median net-
works were constructed using NETWORK, version
4.6.1.0 (http://www.fluxus-engineering.com; Bandelt,
Forster & Rohl, 1999). The network analyses and
subsequent Bayesian clustering analysis were per-
formed for each locus and for the combined dataset
that included all six loci. MC1R-a and -b were
included in combined data A and combined data B,
respectively.
Population genetic structure was inferred by Bayes-
ian clustering analysis using STRUCTURE, version
2.3. This analysis estimates the number of genetic
clusters (K) based on allelic frequencies of loci in the
dataset and survey proportions of genetic clusters
in the populations. Kwas estimated from 5 000 000
Markov chain Monte Carlo (MCMC) generations,
with data sampled every 10 000 generations after a
burn-in period of 2 000 000 generations. Convergence
of MCMC chains was checked based on the similarity
of estimated log probability values for data [lnP(D)]
and the variance of log likelihood {Var [lnP(D)]}
in an independent analysis. Admixture and allele
frequency-correlated models were assumed in the
analyses. To determine an appropriate K, we used
Evanno’s method in STRUCTURE HARVESTER
(Earl & vonHoldt, 2012).
RESULTS
PHYLOGENETIC ANALYSIS OF SRY SEQUENCES
In total, 1002 bp from the SRY region were obtained
from 30 samples (see Supporting information,
Table S2), except for a simple tandem repeat of (GT)n
that was approximately 36 bp in length. The dataset
comprised four haplotypes with six variable sites,
including one base indel. The four haplotypes were
divided into two major lineages (N and S) by five of
POPULATION GENETIC STRUCTURE IN JAPANESE HARE 767
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
the six variable sites. The two lineages showed a
clear allopatric distribution with the border located
approximately halfway between Kanto and Chubu
districts (Fig. 2). The bootstrap value of lineage S
(95%) was substantially high, although the value for
lineage N (62%) was not. Lineage N was composed
of two haplotypes (Haps N1 and N2) from 13 samples
from populations 1, 2, and 4. For lineage S, two
haplotypes (Haps S1 and S2) were found from 17
samples from populations 3 and 5–11. The divergence
time between the two lineages was estimated to
be 1.07 Mya, with 95% highest-probability density
ranging from 0.95 to 2.99 Mya.
GENETIC VARIATIONS IN SIX NUCLEAR
GENE SEQUENCES
For six nuclear loci, 104 samples were used in the
subsequent analyses, including missing loci for which
amplification failed. The sequence lengths for the
six nuclear genes ranged from 316 bp for SPTBN1
to 668 bp for TSHB (see Supporting information,
Table S3; accession number in Table S2). Because the
MC1R gene was estimated to have a recombination
point halfway along the fragment, we partitioned
the sequences into the first 456 bp (MC1R-a) and the
last 375 bp (MC1R-b) and used them separately in
the analyses. The number of polymorphic sites ranged
from two (TYR) to nine (ASIP). The number of alleles
ranged from four (TYR)to12(ASIP). Although MC1R
had five nonsynonymous substitutions at V95I,
A101V, V154M, A179T, and V199I (site numbers
for substitutions refer to the MC1R sequence for
O. cuniculus), in 11 substitutions within sequences
that totalled 831 bp, the geographical distributions
of the variations were restricted to a few populations
and did not appear to match the distributions of
the two morphotypes (see Supporting information,
Table S4). The significance of IBD varied among
genes, and the r-values of the Mantel tests were
generally very small (see Supporting information,
Table S3). SPTBN1 and TSHB exhibited lower
genetic diversity than the other genes, especially in
northern populations. In addition, TSHB had a sig-
nificantly negative Tajima’s D.
GENETIC DIFFERENCES AMONG POPULATIONS BASED
ON SIX NUCLEAR GENES
For the exact test, we also examined allelic differ-
ences between groups that represented the two SRY
lineages discovered in the present study (Table 2).
Almost all genes exhibited significant differences
between the two winter coat groups, between the two
CYT B lineage groups, and between the two SRY
lineage groups (P<0.05). TYR and MC1R did not
differ significantly between Lba and Lbb.
Population networks were constructed using the NJ
method (Saitou & Nei, 1987), pairwise FST distances,
and the significance levels of FST values for the six
autosomal genes (Fig. 3). Here, we describe networks
based on pairwise FST distances (Fig. 3A) and the
significance of FST (Fig. 3B) as ‘distance-based’ and
‘significance-based’ networks, respectively. The rela-
tionships among populations estimated by the two
types of network were generally similar. For example,
population 5 was placed at one end of both TYR
networks. Populations 4, 7, 2, 6, 8, 10, and 11 pro-
ceeded in sequence from the other end of the distance-
based network for TYR (Fig. 3A) and were clu-
stered together in one node in the significance-based
network (Fig. 3B). However, some points differed
between the two types of network for TYR. Popu-
lation 9 was closely related to population 4 in the
Table 2. Summary of nondifferentiation exact Pvalues between two putative groups based on winter coat colour: CYT
B(previous study) and the results for SRY
Gene fragment
Exact Pvalue
Colour (Lba
versus Lbb)
North versus
South of CYT B
North versus
South of SRY
ASIP <0.01 0.01 0.01
TYR 0.13 0.06 0.43
MC1R-a 0.08 0.06 0.01
MC1R-b 0.23 <0.01 <0.01
MC1R (full length) 0.11 <0.01 <0.01
APOB <0.01 <0.01 <0.01
SPTBN1 <0.01 <0.01 <0.01
TSHB <0.01 <0.01 0.03
Significant Pvalues are shown in bold.
768 M. NUNOME ET AL.
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
distance-based network, whereas it was assigned to
the node of populations 1 and 3 and was separated
from population 4 in the significance-based network.
In the networks for ASIP, population 3 was located
near population 11 in the distance-based network,
whereas the significance-based network placed popu-
lation 3 at the same node as population 9, which was
very far from population 11. Similar partial discrep-
ancies were observed between networks for APOB,
in which the relationship between populations 2 and
10 differed. Remarkable differences were found
between networks for combined data B, which con-
sisted of MC1R-b and the other five genes (without
MC1R-a). The significance-based network showed
0.1
pop8
pop7
pop10
pop9
pop11
pop6
pop4
pop3
pop1
pop2
pop5
0.1
pop4
pop1
pop2
pop3
pop5
pop6
pop8
pop7
pop10
pop9
pop11
0.1
pop6
pop9
pop4
pop1
pop2
pop8
pop3
pop11
pop7
pop5
pop10
0.1
pop4
pop9
pop1
pop3
pop5
pop7
pop2
pop6
pop8
pop10, 11
0.1
pop2
pop4
pop5
pop1
pop8
pop11
pop9
pop6
pop3, 7, 10
0.1
pop3
pop10
pop2
pop9
pop11
pop8
pop7
pop5
pop1
pop4
pop6
0.1
pop6
pop8
pop11
pop7, 9
pop10
pop5
pop2, 3, 4 pop1
0.1
pop7
pop2
pop5
pop1
pop6
pop11
pop10
pop3
pop4
pop8
pop9
0.1
pop7
pop11
pop2
p
op1, 4
pop8
pop10
pop6
pop9
pop3
pop5
ASIP TYR MC1R-a
MC1R-b APOB SPTBN1
TSHB Combined (MC1R-a) Combined (MC1R-b)
A
Figure 3. Population networks for six nuclear genes and combined data for all loci. A, networks calculated based on
pairwise FST distance. B, networks constructed based on significance of FST. Numbers on branches of networks in (B) are
population numbers and indicate the cut-off point for significant differences for a certain population. Combined data A
includes all sequences except MC1R-b (the latter half of MC1R) and combined data B includes all sequences except
MC1R-a (the first half of MC1R).
POPULATION GENETIC STRUCTURE IN JAPANESE HARE 769
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
close connections among populations 1, 2, 9, and 11,
despite large genetic and geographical distances
among these populations in the distance-based
network. Although several differences were observed
between the two types of population network, popu-
lations 1, 2, 9, and 11 tended to be located at external
nodes of the networks for all six genes. Combined
data A and B subdivided the other populations (popu-
lations 3–8 and 10) into two similar groups: popula-
tions 3–6 and populations 7, 8, and 10.
Bayesian clustering analysis indicated that the
Japanese hare population is composed of four genetic
clusters (N1 and S1–S3), the geographical distri-
butions of which show a north–south cline (Fig. 4).
STRUCTURE HARVESTER indicated that K= 4 was
the most appropriate parameter for combined data A
and B. Because these datasets produced very similar
results, we used only the results from combined data A.
FST genetic distances among the four clusters indicated
that cluster N was very distant from the other three.
B
pop6
pop4
pop1
pop2
pop8
p
op3, 9
pop11
pop7
pop5
pop10
9
3
41
2
1
2
7
11
4
4
8
8
8
8
10
10
2
211
11
7
ASIP
pop1, 3, 9
pop5
pop2, 4, 6, 7, 8, 10, 11
5
11
10
8
7
6
4
2
pop1
11
pop2
pop8
pop4, 5, 9
pop6, 11
pop7, 10
6
10
7
3
1
2
8
pop3
pop9
7
pop6
pop2, 3, 7, 8, 10
pop11
8
10
11
3
1
2
8
pop1, 4
2
3
pop5
pop9
8
pop2
11 pop11
pop10
pop4, 6
pop3, 7, 8
pop5
pop1
3
4
7
9
5
11
1
10
3
7
8
2
2
1
11
4
6
6
pop1, 2, 3, 4
6
pop6, 7, 9, 11
pop5, 10
pop8
7
9
11 4
2
3
1
4
2
3
16
7
9
11
5
8
10
pop9
6
pop8
pop3,5,6,7,10
pop4
7
10
4
1
9
11
2
pop1, 11
3
5
pop2
8
4
pop2
pop1
pop8
pop7, 10
pop9, 11
pop6
pop3
pop4, 5
1
6
10
7
9
11
9
11
10
7
2
6
6
6
4
5
4
5
10
7
2
8
6
6
8
8
pop4, 5
pop3 pop6
pop2
pop11
pop1
pop9
pop7 pop8
pop10
7
10 2
8
4
56
3
7
10
19
8
3
11
2
2
711
TYR MC1R-a
MC1R-b APOB SPTBN1
TSHB Combined (MC1R-a) Combined (MC1R-b)
Figure 3. Continued.
770 M. NUNOME ET AL.
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
Clusters S2 and S3 branched off from cluster S1, with
a relatively shorter FST distance compared to that
observed between S1 and N. Cluster N prevailed in the
northernmost population (population 1). Populations 2
and 4 also had relatively high frequencies for cluster N
relative to the other populations. The S clusters were
broadly observed in populations 2–11. Cluster S1 gen-
erally included individuals from populations 8 and 11
(Fig. 4A), and clusters S2 and S3 were equally distrib-
uted over all populations, except population 1.
DISCUSSION
USEFULNESS OF A POPULATION NETWORK BASED ON
THE SIGNIFICANCE OF FST VALUES FOR DEPICTING
POPULATION GENETIC STRUCTURE
Population networks based on the (non)significance
of pairwise FST distances proved rather effective for
clustering populations compared to conventional net-
works of FST distance. Significance-based networks
had almost the same topologies as conventional net-
works of FST distance for every gene and combined
data A; in combined data B, populations 1, 2, 9, and
11 were placed near one another (Fig. 3B). Although
populations 1 and 2 and populations 9 and 11 showed
substantial genetic differences from one another,
they were similar in that they also had significant
genetic differences from the other populations in com-
bined data B. Although the distance-based network
could recognize the large genetic differences between
populations 1 and 2 and populations 9 and 11, the
significance-based network treated all significant
genetic differences among populations as having
the same value of ‘1’, which did not reflect genetic
distance. For this reason, distance- and significance-
based networks of combined data B produced differ-
ent topologies. Thus, combined data A provided a
more reasonable significance-based network, in which
the four groups of populations could be recognized
more easily than in the conventional FST network: (1)
populations 1 and 2; (2) populations 3–6; (3) popula-
tions 7, 8, and 10; and (4) populations 9 and 11. The
significance-based network with combined data A
appeared to reflect the results of the population clus-
tering analyses (Fig. 4). For example, the primary
genetic component in populations 1 and 2 was cluster
N1, and populations 3–6 contained the four genetic
clusters with similar proportions. Although simula-
tion studies would be needed to evaluate the efficiency
of the significance-based network, this may provide
an easy method for surveying clusters of populations
that showed significant genetic differences.
TWO DIVERGENT GENETIC GROUPS IN
L. BRACHYURUS AS A RESULT OF PAST VICARIANCE
The SRY and nuclear genes showed two genetically
divergent groups in L. brachyurus, as demonstrated
pop. 1 pop. 2 pop. 3 pop. 11pop. 10pop. 9pop. 8pop. 7pop. 6pop. 5pop. 4
N1S1
S2
S3
0.01
Cluster N1
Cluster S1
Cluster S2
Cluster S3
A
B
Figure 4. Results of genetic structuring analyses using STRUCTURE, version 2.3. A, genetic component proportions of
individuals (upper) and populations (lower) are represented by thin horizontal bars composed of four coloured clusters.
The component proportions of a cluster indicate the proportions of assignment of the individual (population) into each of
the four subgroups. B, genetic relationships among clusters are illustrated using networks based on FST genetic distance.
POPULATION GENETIC STRUCTURE IN JAPANESE HARE 771
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
previously using mitochondrial data (Nunome et al.,
2010), even though genetic differences between the
two morphotypes were obscure in our molecular data
(Figs 2, 4). A geographical border between the two
divergent lineages in SRY was clearly located across
the eastern part of Honshu Island, running from
north to south, although the location did not match
the distribution of either morphotype. Interestingly,
the border did not even match the two CYT B lineages
in Chugoku district on the western part of Honshu
Island (Fig. 2). The presence of two lineages within
the species was well supported by nuclear genes, with
two divergent genetic clusters (N and S) alternately
in the northern and southern areas (Fig. 4). When
STRUCTURE analysis was performed on the basis
of the K= 2 model, clusters S1, S2, and S3 were
combined into one cluster ‘S’ and then only two clus-
ters ‘N’ and ‘S’ were exhibited without any changes
in frequency in populations (data not shown). The
existence of two SRY lineages and the presence of
two genetic clusters in the nuclear genes could be
explained by past vicariance in the species, as hypoth-
esized in a previous mitochondrial study (Nunome
et al., 2010). In addition, the timing of the vicariance
in the two SRY lineages was estimated to be approxi-
mately 1 Mya, which is similar to the estimate of
1.2 Mya for the two CYT B lineages, although the
evolutionary rate of the SRY sequences used in the
present study would have been slow, given the small
numbers of substitutions observed in the sequences.
Genetic subdivisions in morphology, karyotype, and
DNA sequences have been found commonly at inter-
or intraspecific levels in various Japanese mammals
(Nagata et al., 1999; Tsuchiya et al., 2000; Iwasa &
Abe, 2006; Kawamoto et al., 2007; Tomozawa &
Suzuki, 2008; Oshida, Masuda & Ikeda, 2009; Yasuda
et al., 2012). One possible explanation for these diver-
gences is repeated habitat fragmentation during
glacial periods in the Late Pleistocene. The four SRY
haplotypes (Fig. 2) and three subclusters of cluster S
(S1–S3; Fig. 4) would also be products of fragmenta-
tion during a more recent period. Although we could
not determine a relationship between past vicariance
and morphological divergence in L. brachyurus,a
vestige of the vicariance would remain in the strong
genetic population structure.
NUCLEAR GENES IMPLY A GENETIC MIXTURE
BETWEEN TWO MORPHOTYPES
The results of the present study suggest that Lba
and Lbb were genetically indistinguishable from each
other, in contrast to our initial expectation that the
two morphotypes with different winter coat colours
would be genetically differentiated as a result of dif-
ferent evolutionary trajectories. The topologies of the
population networks were not consistent among the
six genes. The absence of consistency among the
networks for the six genes implies that the popula-
tions are not genetically structured, as is evident in
the mixing states of the four genetic clusters in the
nuclear genes among the populations. Similar findings
were reported in willow grouse and Arctic fox, which
also showed little genetic divergence between popula-
tions with different winter coat colours (Meinke et al.,
2001; Skoglund & Hoglund, 2010). Genetic variation
at neutral loci that are not correlated with a colour
polymorphism or genes involved in a polymorphism
have been commonly observed in other mammals,
such as the oldfield mouse (Peromyscus polionotus)
(Mullen & Hoekstra, 2008), the rock pocket mouse
(Chaetodipus intermedius) (Hoekstra, Drumm &
Nachman, 2004), and the grey wolf (Canis lupus)
(Anderson et al., 2009). The two morphotypes of
L. brachyurus may be only differentiated at genes
involved in the differences in seasonal changes in
coat colour. Thus, genome-wide surveys using abun-
dant genetic markers, such as microsatellites or
single-nucleotide polymorphism analyses, in individu-
als around the boundary between Lba and Lbb are
required to fully appreciate gene flow between the two
morphotypes and to determine which loci show affini-
ties with the morphotypes.
GENETIC VARIATION OF COAT COLOUR-RELATED
GENES AND NEUTRAL GENES AMONG POPULATIONS
Initially, we expected that genes involved in differ-
ences in winter coat colour would show particular
geographical differences in genetic diversity. None of
the markers, including the three coat colour-related
genes, exhibited genetic differentiation between the
two morphotypes. Tajima’s neutrality tests did not
suggest that natural selection was affecting the genes
(Fig. 3; see also Supporting information, Table S3).
No differences in genetic diversity (HEor pi) among
populations were observed in the three genes. The
results did not suggest that environmental differences
between locations were affecting the genes. We found
five amino acid variants of MC1R (V154M, A101V,
V95I, V199I, and A179T) (see Supporting informa-
tion, Table S4). One of the variants, A101V, is placed
in the region where the jaguar (Panthera onca) has
a 15-bp deletion related to its coat colour variation
(Eizirik et al., 2003). The remaining four amino acid
variants did not correspond to any of the amino acid
changes related to coat colour variations of other
animals (Majerus & Mundy, 2003). However, all five
mutations appeared to be distributed without any
relation to the areas of Lba and Lbb. Similar findings
were reported in the Japanese marten (Hosoda
et al., 2005; Sato et al., 2009) and the willow grouse
772 M. NUNOME ET AL.
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
(Skoglund & Hoglund, 2010). MC1R and other
neutral markers showed no clear genetic differences
between the two morphotypes in the Japanese
marten. In addition, genetic variation in TYR was not
related to differences in winter colour in the willow
grouse. Although a study of the Arctic fox found a
relationship between substitutions in MC1R and vari-
ation in winter coat colour (Vage et al., 2005), MC1R,
as well as ASIP and TYR, are involved in lifetime
coat-colour variations in other mammals (Klungland
& Vage, 2003; Nachman et al., 2003; Fontanesi et al.,
2006; Kambe et al., 2011). Thus, the seasonal change
in coat colour in L. brachyurus may be controlled by
other regulatory genes and not by melanogenesis-
related genes such as MC1R and ASIP.
The negative value of Tajima’s Dfor TSHB and
its lower genetic variation in northern populations
suggested that the locus is under natural selection.
Although TSHB has been used as a neutral marker
for phylogenetic analyses (Matthee et al., 2004;
Hoofer et al., 2008; Stoffberg et al., 2010), it has
recently been reported to regulate seasonal reproduc-
tion through photoperiodic signalling in birds and
mammals (Nakao et al., 2008; Ono et al., 2008). Wild
stickleback populations showed differences in gene
expression for TSHβ2, a paralogue of TSHB, between
marine and stream ecotypes, especially under short-
photoperiod conditions (Kitano et al., 2010). Genetic
variation in photoperiod-related genes could vary
with latitude, as was shown in European populations
of Drosophila melanogaster (Tauber et al., 2007)
and Alaskan populations of Chinook salmon
(Oncorhynchus tshawytscha) (O’Malley & Banks,
2008). Because L.brachyurus is known to show
seasonal changes in reproductive activity with
photoperiodic signals (Otsu, 1971), TSHB could be
subject to the latitudinal cline in seasonal photo-
period in the Japanese archipelago. To further
examine this issue, studies of the seasonal and geo-
graphical expression patterns of the gene in wild
mammals that show seasonal changes in breeding
activity or coat colour, such as Lepus and Martes
species, are needed.
CONCLUSIONS
The results of the present study suggest that genetic
diversity in the Japanese hare is attributable to
vicariance events, which may have occurred approxi-
mately 1 Mya, rather than to differences in winter
coat colour. Although the two morphotypes of Japa-
nese hare have been considered to be different sub-
species, the difference in winter coat colour did not
appear to restrain gene flow. However, some func-
tional genes related to seasonal changes in coat colour
and reproductive status may vary geographically
according to environmental differences within the
Japanese archipelago. In this case, we can carry out
population genetic surveys of the Japanese hare
across its range because the Japanese hare inhabits a
restricted area: the Japanese archipelago. In addition,
the accumulation of genomic information for rabbit
(O. cuniculus) would facilitate genome-wide research
on the Japanese hare, although these two species are
not closely related. Indeed, several primers used in
the present study were designed based on the pub-
lished genome sequences of the rabbit. Many rabbit
microsatellite markers have been applied for popula-
tion genetic studies of Lepus species (Hamill, Doyle &
Duke, 2006; Thulin, Fang & Averianov, 2006) and
a high karyotype similarity was revealed between
Lepus and Oryctolagus (Robinson, Yang & Harrison,
2002). Thus, to detect a candidate gene for seasonal
changes in coat colour and to survey gene flow
between populations that show differences in winter
coat colour, the Japanese hare represents an appro-
priate model species.
ACKNOWLEDGEMENTS
We thank Kimiyuki Tsuchiya, Shimane Prefecture
Mountainous Region Research Centre and Toyama
Family Park, for providing tissue specimens. We also
express our appreciation to our colleagues for their
valuable advice during the present study. This study
was supported in part by a Grant-in-Aid for Fellows
and for Research Activity Start-Up of the Japan
Society for the Promotion of Science from the Minis-
try of Education, Culture, Sports, Science and Tech-
nology, Japan.
REFERENCES
Aigner B, Besenfelder U, Muller M, Brem G. 2000.
Tyrosinase gene variants in different rabbit strains. Mam-
malian Genome 11: 700–702.
Alves PC, Melo-Ferreira J, Freitas H, Boursot P. 2008.
The ubiquitous mountain hare mitochondria: multiple
introgressive hybridization in hares, genus Lepus.Philo-
sophical Transactions of the Royal Society B, Biological
Sciences 363: 2831–2839.
Amrine-Madsen H, Koepfli KP, Wayne RK, Springer MS.
2003. A new phylogenetic marker, apolipoprotein B, pro-
vides compelling evidence for eutherian relationships.
Molecular Phylogenetics and Evolution 28: 225–240.
Anderson TM, vonHoldt BM, Candille SI, Musiani M,
Greco C, Stahler DR, Smith DW, Padhukasahasram B,
Randi E, Leonard JA, Bustamante CD, Ostrander EA,
Tang H, Wayne RK, Barsh GS. 2009. Molecular and
evolutionary history of melanism in North American gray
wolves. Science 323: 1339–1343.
POPULATION GENETIC STRUCTURE IN JAPANESE HARE 773
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
Bandelt HJ, Forster P, Rohl A. 1999. Median-joining net-
works for inferring intraspecific phylogenies. Molecular
Biology and Evolution 16: 37–48.
Chen Y, Duhl DM, Barsh GS. 1996. Opposite orientations of
an inverted duplication and allelic variation at the mouse
agouti locus. Genetics 144: 265–277.
Drummond A, Rambaut A, Shapiro B, Pybus O. 2005.
Bayesian coalescent inference of past population dynamics
from molecular sequences. Molecular Biology and Evolution
22: 1185–1192.
Earl DA, vonHoldt BM. 2012. STRUCTURE HARVESTER:
a website and program for visualizing STRUCTURE output
and implementing the Evanno method. Conservation Genet-
ics Resources 4: 359–361.
Eizirik E, Yuhki N, Johnson WE, Menotti-Raymond M,
Hannah SS, O’Brien SJ. 2003. Molecular genetics and
evolution of melanism in the cat family. Current Biology 13:
448–453.
Excoffier L, Lischer HEL. 2010. Arlequin suite ver 3.5: a new
series of programs to perform population genetics analyses
under Linux and Windows. Molecular Ecology Resources 10:
564–567.
Filatov D. 2001. Processor of sequences manual. University of
Birmingham. Available at: http://www.biosciences.bham.ac
.uk/labs/filatov/proseq.html
Flux JEC. 1970. Colour change of mountain hares (Lepus
timidus scotius) in north-east Scotland. Zoology 162: 345–
358.
Fontanesi L, Tazzoli M, Beretti F, Russo V. 2006.
Mutations in the melanocortin 1 receptor (MC1R) gene
are associated with coat colors in the domestic rabbit
(Oryctolagus cuniculus). Animal Genetics 37: 489–493.
Grange WB. 1932. The pelages and color changes of the
snowshoe hare, Lepus americanns phaeonotus Allen.
Journal of Mammalogy 13: 99–116.
Guindon S, Gascuel O. 2003. A simple, fast, and accurate
algorithm to estimate large phylogenies by maximum like-
lihood. Systematic Biology 52: 696–704.
Guindon S, Lethiec F, Duroux P, Gascuel O. 2005.
PHYML online – a web server for fast maximum likelihood-
based phylogenetic inference. Nucleic Acids Research 33:
W557–W559.
Hamill RM, Doyle D, Duke EJ. 2006. Spatial patterns
of genetic diversity across European subspecies of the
mountain hare, Lepus timidus L. Heredity 97: 355–
365.
Hewson R. 1958. Moults and winter whitening in the Moun-
tain hare Lepus timidus scoticus Hilzheimer. Proceedings of
the Zoological Society of London 131: 99–108.
Hirata S. 1999. Nousagi no hanashi. Akita: Mumyosha
Shuppan (in Japanese).
Hoekstra HE, Drumm KE, Nachman MW. 2004. Ecological
genetics of adaptive color polymorphism in pocket mice:
geographic variation in selected and neutral genes. Evolu-
tion 58: 1329–1341.
Hoofer SR, Flanary WE, Bull RJ, Baker RJ. 2008.
Phylogenetic relationships of vampyressine bats and allies
(Phyllostomidae: Stenodermatinae) based on DNA sequences
of a nuclear intron (TSHB-I2). Molecular Phylogenetics and
Evolution 47: 870–876.
Hosoda T, Sato JJ, Shimada T, Campbell KL, Suzuki H.
2005. Independent non-frameshift deletions in the MC1R
gene are not associated with melanistic coat coloration
in three mustelid lineages. Journal of Heredity 95: 607–
623.
Hudson R, Kaplan N. 1985. Statistical properties of the
number of recombination events in the history of a sample
of DNA sequences. Genetics 111: 147–164.
Iason GR, Ebling JP. 1989. Seasonal variation in the daily
pattern of plasma melatonin in a wild mammal: the moun-
tain hare (Lepus timidus). Journal of Pineal Research 6:
157–167.
Imaizumi Y. 1960. Coloured illustrations of the mammals of
Japan. Osaka: Hoikusha (in Japanese).
Iwasa MA, Abe H. 2006. Colonization history of the Japa-
nese water shrew Chimarrogale platycephala, in the Japa-
nese Islands. Acta Theriologica 51: 29–38.
Kalinowski ST. 2009. How well do evolutionary trees
describe genetic relationships between populations? Hered-
ity 102: 506–513.
Kambe Y, Tanikawa T, Matsumoto Y, Tomozawa M,
Aplin KP, Suzuki H. 2011. Origin of agouti-melanistic
polymorphism in wild black rats (Rattus rattus) inferred
from Mc1r gene sequences. Zoological Science 28: 560–567.
Kawamoto Y, Shotake T, Nozawa K, Kawamoto S,
Tomari K, Kawai S, Shirai K, Morimitsu Y, Takagi N,
Akaza H, Fujii H, Hagihara K, Aizawa K, Akachi S, Oi
T, Hayaishi S. 2007. Postglacial population expansion
of Japanese macaques (Macaca fuscata) inferred from
mitochondrial DNA phylogeography. Primates 48: 27–40.
Kinoshita G, Nunome M, Han SH, Hirakawa H,
Suzuki H. 2012. Ancient colonization and within-island
vicariance revealed by mitochondrial DNA phylogeography
of the mountain hare (Lepus timidus) in Hokkaido, Japan.
Zoological Science 29: 776–785.
Kitano J, Lema SC, Luckenbach JA, Mori S, Kawagishi
Y, Kusakabe M, Swanson P, Peichel CL. 2010. Adaptive
divergence in the thyroid hormone signaling pathway in the
stickleback radiation. Current Biology 20: 2124–2130.
Klungland H, Vage DI. 2003. Pigmentary switches in
domestic animal species. Annals of the New York Academy of
Sciences 994: 331–338.
Koutsogiannouli EA, Moutou KA, Stamatis C,
Suchentrunk F, Mamuris Z. 2012. Analysis of MC1R
genetic variation in Lepus species in Mediterranean refugia.
Mammalian Biology 77: 428–433.
Kuderling I, Cedrini MC, Fraschini F, Spagnesi M. 1984.
Season-dependent effects of melatonin on testes and fur
colour in mountain hares (Lepus timidus L.). Experientia 40:
501–502.
Librado P, Rozas J. 2009. DnaSP v5: a software for com-
prehensive analysis of DNA polymorphism data. Bioinfor-
matics 25: 1451–1452.
Liu J, Yu L, Arnold ML, Wu CH, Wu SF, Lu X, Zhang YP.
2011. Reticulate evolution: frequent introgressive hybridi-
zation among chinese hares (genus Lepus) revealed by
774 M. NUNOME ET AL.
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
analyses of multiple mitochondrial and nuclear DNA loci.
BMC Evolutionary Biology 11: 223.
Majerus MEN, Mundy NI. 2003. Mammalian melanism:
natural selection in black and white. Trends in Genetics 19:
585–588.
Matthee C, van Vuuren B, Bell D, Robinson T. 2004.
A molecular supermatrix of the rabbits and hares (Lepori-
dae) allows for the identification of five intercontinental
exchanges during the Miocene. Systematic Biology 53: 433–
447.
Meinke PG, Kapel CMO, Arctander P. 2001. Genetic dif-
ferentiation of populations of Greenlandic Arctic fox. Polar
Research 20: 75–83.
Miller M. 2005. Alleles In Space (AIS): computer software for
the joint analysis of interindividual spatial and genetic
information. Journal of Heredity 96: 722–724.
Mills LS, Zimova M, Oyler J, Running S, Abatzoglou JT,
Lukacs PM. 2013. Camouflage mismatch in seasonal
coat color due to decreased snow duration. Proceedings of
the National Academy of Sciences of the United States of
America 110: 7360–7365.
Miltenberger RJ, Wakamatsu K, Ito S, Woychik RP,
Russell LB, Michaud EJ. 2002. Molecular and phenotypic
analysis of 25 recessive, homozygous-viable alleles at the
mouse agouti locus. Genetics 160: 659–674.
Mullen LM, Hoekstra HE. 2008. Natural selection along an
environmental gradient: a classic cline in mouse pigmenta-
tion. Evolution 62: 1555–1570.
Nachman MW, Hoekstra HE, D’Agostino SL. 2003. The
genetic basis of adaptive melanism in pocket mice. Proceed-
ings of the National Academy of Sciences of the United
States of America 100: 5268–5273.
Nagata J, Masuda R, Tamate HB, Hamasaki S,
Ochiai K, Asada M, Tatsuzawa S, Suda K, Tado H,
Yoshida MC. 1999. Two genetically distinct lineages of
the sika deer, Cervus nippon, in Japanese islands: compari-
son of mitochondrial D-loop region sequences. Molecular
Phylogenetics and Evolution 13: 511–519.
Nakao N, Ono H, Yamamura T, Anraku T, Takagi T,
Higashi K, Yasuo S, Katou Y, Kageyama S, Uno Y,
Kasukawa T, Iigo M, Sharp PJ, Iwasawa A, Suzuki Y,
Sugano S, Niimi T, Mizutani M, Namikawa T, Ebihara
S, Ueda HR, Yoshimura T. 2008. Thyrotrophin in the pars
tuberalis triggers photoperiodic response. Nature 452: 317–
322.
Nei M, Li WH. 1979. Mathematical model for studying
genetic variation in terms of restriction endonucleases. Pro-
ceedings of the National Academy of Sciences of the United
States of America 76: 5269–5273.
Nunome M, Torii H, Matsuki R, Kinoshita G, Suzuki H.
2010. The influence of pleistocene refugia on the evolution-
ary history of the Japanese hare, Lepus brachyurus.Zoo-
logical Science 27: 746–754.
O’Malley KG, Banks MA. 2008. A latitudinal cline in the
Chinook salmon Oncorhynchus tshawytscha Clock gene: evi-
dence for selection on PolyQ length variants. Proceedings
of the Royal Society of London B, Biological Sciences 275:
2813–2821.
Ono H, Hoshino Y, Yasuo S, Watanabe M, Nakane Y,
Murai A, Ebihara S, Korf HW, Yoshimura T.
2008. Involvement of thyrotropin in photoperiodic signal
transduction in mice. Proceedings of the National Academy
of Sciences of the United States of America 105: 18238–18242.
Oshida T, Masuda R, Ikeda K. 2009. Phylogeography of
the Japanese giant flying squirrel, Petaurista leucogenys
(Rodentia: Sciuridae): implication of glacial refugia in an
arboreal small mammal in the Japanese Islands. Biological
Journal of the Linnean Society 98: 47–60.
Otsu S. 1967. Ecological studies of Tohoku hare, Lepus
brachyurus angustidens Hollister. III. On factors controlling
coat color change. Japanese Journal of Applied Entomology
and Zoology 11: 37–42 (in Japanese).
Otsu S. 1971. Ecological studies of Tohoku hare, Lepus
brachyurus angustidens Hollister IV. Effect of prolonged
exposure to light on breeding. Japanese Journal of Applied
Entomology and Zoology 15: 31–35 (in Japanese).
Posada D, Crandall K. 1998. MODELTEST: testing
the model of DNA substitution. Bioinformatics 14: 817–
818.
Pritchard JK, Stephens M, Donnelly P. 2000. Inference of
population structure using multilocus genotype data. Genet-
ics 155: 945–959.
Rambaut A, Drummond AJ. 2007. Tracer, Version 1.4.
Available at: http://beast.bio.ed.ac.uk/Tracer
Raymond M, Rousset F. 1995. An exact test for population
differentiation. Evolution 49: 1280–1283.
Robinson TJ, Yang F, Harrison WR. 2002. Chromosome
painting refines the history of genome evolution in hares
and rabbits (order Lagomorpha). Cytogenetic and Genome
Research 96: 223–227.
Russell JE, Tumlison R. 1996. Comparison of microstruc-
ture of white winter fur and brown summer fur of some
Arctic mammals. Acta Zoologica 77: 279–282.
Rust CC. 1965. Hormonal control of pelage cycles in the
short-tailed weasel (Mustela erminea bangsi). General and
Comparative Endocrinology 5: 222–231.
Saitou N, Nei M. 1987. The neighbor-joining method: a new
method for reconstructing phylogenetic trees. Molecular
Biology and Evolution 4: 406–425.
Sambrook J, Russell DW. 2001. Molecular cloning. A labo-
ratory manual, 3rd edn. New York, NY: Cold Spring Harbor
Laboratory Press.
Sato JJ, Yasuda SP, Hosoda T. 2009. Genetic diversity of
the Japanese marten (Martes melampus) and its implica-
tions for the conservation unit. Zoological Science 26: 457–
466.
Scherbarth F, Steinlechner S. 2010. Endocrine mecha-
nisms of seasonal adaptation in small mammals: from early
results to present understanding. Journal of Comparative
Physiology B 180: 935–952.
Skoglund P, Hoglund J. 2010. Sequence polymorphism in
candidate genes for differences in winter plumage between
Scottish and Scandinavian willow grouse (Lagopus lagopus).
PLoS One 5: e10334.
Stoffberg SD, Jacobs S, MacKie IJ, Matthee CA. 2010.
Molecular phylogenetics and historical biogeography of
POPULATION GENETIC STRUCTURE IN JAPANESE HARE 775
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
Rhinolophus bats. Molecular Phylogenetics and Evolution
54: 1–9.
Stoner CJ, Bininda-Emonds ORP, Caro T. 2003. The
adaptive significance of coloration in lagomorphs. Biological
Journal of the Linnean Society 79: 309–328.
Swofford DL. 2002. PAUP*. Phylogenetic analysis using par-
simony (*and other methods), Version 4. Sunderland, MA:
Sinauer Associates.
Tajima F. 1989. Statistical method for testing the neutral
mutation hypothesis by DNA polymorphism. Genetics 123:
585–595.
Tauber E, Zordan M, Sandrelli F, Pegoraro M,
Osterwalder N, Breda C, Daga A, Selmin A, Monger K,
Benna C, Rosato E, Kyriacou CP, Costa R. 2007.
Natural selection favors a newly derived timeless allele in
Drosophila melanogaster.Science 316: 1895–1898.
Thulin CG, Fang MY, Averianov AO. 2006. Introgression
from Lepus europaeus to L-timidus in Russia revealed by
mitochondrial single nucleotide polymorphisms and nuclear
microsatellites. Hereditas 143: 68–76.
Tomozawa M, Suzuki H. 2008. A trend of central versus
peripheral structuring in mitochondrial and nuclear gene
sequences of the Japanese wood mouse, Apodemus speciosus.
Zoological Science 25: 273–285.
Tsuchiya K, Suzuki H, Shinohara A, Harada M, Wakana
S, Sakaizumi M, Han S, Lin L, Kryukov A. 2000.
Molecular phylogeny of East Asian moles inferred from the
sequence variation of the mitochondrial cytochrome bgene.
Genes and Genetic Systems 75: 17–24.
Vage DI, Fuglei E, Snipstad K, Beheim J, Landsem VM,
Klungland H. 2005. Two cysteine substitutions in the
MC1R generate the blue variant of the arctic fox (Alopex
lagopus) and prevent expression of the white winter coat.
Peptides 26: 1814–1817.
Walsberg GE. 1991. Thermal effects of seasonal coat change
in 3 sub-arctic mammals. Journal of Thermal Biology 16:
291–296.
Watson A. 1963. The effect of climate on the colour of moun-
tain hares in Scotland. Proceedings of the Zoological Society
of London 41: 823–835.
Watson A. 1973. Moults of wild Scottish ptarmigan, Lagopus
mutus, in relation to sex, climate and status. Journal of
Zoology 171: 207–223.
Wright S. 1951. The genetical structure of population. Annals
of Eugenics 15: 323–354.
Wu C, Wu J, Bunch T, Li Q, Wang Y, Zhang Y. 2005.
Molecular phylogenetics and biogeography of Lepus in
Eastern Asia based on mitochondrial DNA sequences.
Molecular Phylogenetics and Evolution 37: 45–61.
Yamada F, Takaki M, Suzuki H. 2002. Molecular phylogeny
of Japanese Leporidae, the Amami rabbit Pentalagus
furnessi, the Japanese hare Lepus brachyurus, and the
mountain hare Lepus timidus, inferred from mitochondrial
DNA sequences. Genes and Genetic Systems 77: 107–116.
Yasuda SP, Iwabuchi M, Aiba H, Minato S, Mitsuishi K,
Tsuchiya K, Suzuki H. 2012. Spatial framework of
nine distinct local populations of the Japanese dormouse
Glirulus japonicus based on matrilineal cytochrome band
patrilineal SRY gene sequences. Zoological Science 29: 111–
120.
SUPPORTING INFORMATION
Additional Supporting Information may be found in the online version of this article at the publisher’s web-site:
Table S1. Primer information.
Table S2. List of accession numbers of samples. Alleles of heterozygous individuals have two accession
numbers. Samples that could not be amplified are represented by hyphens.
Table S3. Molecular indices for loci.
Table S4. Geographical distribution of nonsynonymous substitutions of MC1R.
776 M. NUNOME ET AL.
© 2014 The Linnean Society of London, Biological Journal of the Linnean Society, 2014, 111, 761–776
... These studies suggest that frequent and massive introgressions of mtDNA among East Asian hares have led to confusion regarding their taxonomic and phylogenetic relationships. On the other hand, L. brachyurus, the hare species endemic to the Japanese Archipelago, is suggested to have a long independent history, beginning before the Middle Pleistocene (Nunome et al., 2010(Nunome et al., , 2014Yamada et al., 2002), despite the possibility of secondary contact and introgressive hybridization with L. timidus and other continental species across land bridges during glacial periods. Comparative genetic studies on continents and insular isolates are currently of fundamental importance to understand the geographical factors in evolutionary divergence and shaping biodiversity of a focal area (Patiño et al., 2017), while only a handful of phylogeographic studies have been conducted in East Asia including both sides of the archipelago and the continent (e.g., Aoki et al., 2018;Kinoshita et al., 2015;Sakka et al., 2010). ...
... In total, 97 tissue samples from four Lepus species-L. timidus (n = 42), L. mandshuricus (n = 14), L. coreanus (n = 10), and L. brachyurus (n = 31)-were obtained from road-killed or hunted animals that were collected for the present study and previous studies (Kinoshita et al., 2012;Nunome et al., 2010Nunome et al., , 2014. Sampling locations are shown in Fig. 1 and Supplementary Table 1. ...
... Sequences of the two nDNA loci (Mgf and Phka2) represented intron regions; the Mc1r sequence was from an exon region; and Tg, Tshb, Sptbn1, Asip, and Sry sequences included both intron and exon regions. The amplified region for each locus and the primers used are shown in Supplementary Table 2. Sequences of the Cytb and Sry loci for some samples derived from our previous studies (Kinoshita et al., 2012;Nunome et al., 2010Nunome et al., , 2014 were downloaded from international DNA databases. Novel sequences used in the present study were registered in databases under the accession numbers LC131889-LC132693. ...
... Therefore, the JP-CB population may have experienced further geographic isolation with respect to JP-CCB and JP-YD populations, thus accumulating distinctive genetic variation (note the much longer branch of JP-CB in the phylogeny of Fig. 2B). Other studies with Japanese native species also found substantial genetic divergence between northern and central regions in Honshu, including plants (Senni et al., 2005;Ikeda et al., 2006;Hiraoka & Tomaru, 2009;Qiu et al., 2009), insects (Sota & Hayashi, 2007;Schoville et al., 2013;Saito & Tojo, 2016), and vertebrates (Oshida et al., 2009;Setiamarga et al., 2009;Nunome et al., 2010). ...
Article
Full-text available
Disjunct distribution is a key issue in biogeography and ecology, but it is often difficult to determine relative roles of dispersal vs. vicariance in disjunctions. We studied phylogeographic pattern of the monotypic Conandron ramondioides (Gesneriaceae), which shows Sino-Japanese disjunctions, with ddRAD sequencing based on a comprehensive sampling of 11 populations from mainland China, Taiwan Island, and Japan. We found a very high degree of genetic differentiation among these three regions, with very limited gene flow and a clear Isolation by Distance pattern. Mainland China and Japan clades diverged first from a widespread ancestral population in middle Miocene, followed by a later divergence between mainland China and Taiwan Island clades at early Pliocene. Three current groups have survived in various glacial refugia during Last Glacial Maximum (LGM), and experienced contraction and/or bottlenecks since their divergence during Quaternary glacial cycles, with strong niche divergence between mainland China + Japan and Taiwan Island ranges. Thus, we verified a predominant role of vicariance in the current disjunction of monotypic genus Conandron. The sharp phylogenetic separation, ecological niche divergences among these three groups and the great number of private alleles in all populations sampled indicate a considerable time of independent evolution, and suggests the need of a taxonomic survey to detect potentially overlooked taxa. This article is protected by copyright. All rights reserved.
... These islands are arranged along a north-south axis, and each has a different climate. Numerous species exist in the Japanese archipelago, many of which have colonized the archipelago since millions of years ago (Serizawa et al. 2000;Nunome et al. 2007Nunome et al. , 2010Kirihara et al. 2013;Honda et al. 2019;Nakamoto et al. 2021), allowing comparison of the effects of Quaternary environmental changes within and between species. Several evolutionary studies *To whom correspondence should be addressed. ...
Article
Full-text available
Quaternary environmental change provided opportunities for rapid population expansion; however, the process of building the population spatial structures remains poorly understood. In this study, we determined the mitochondrial cytochrome b and control region sequences of 43 individuals of the large Japanese wood mouse (Apodemus speciosus) from Hokkaido, northern Japan and analyzed these data along with those from 40 other individuals. Consistent with the findings of our previous study, we found that two rapid expansion events, after the last glacial maximum (LGM) and Marine Isotope Stage (MIS) 4, shaped population genetic pattern of A. speciosus in Hokkaido. In northeastern Hokkaido, several ancient lineages that originated during MIS 3 were detected, whereas central Hokkaido was dominated by haplotypes descended from a single lineage that survived the LGM, suggesting that the populations of western part of Hokkaido were newly formed by westward migration from eastern Hokkaido during the post-LGM warm period. Alternatively, as post-LGM vegetation recovery is thought to have occurred gradually from west to east in Hokkaido, population expansion started in the west and moved gradually to the east, resulting in eastward haplotype movement; thus, western and eastern Hokkaido may have served as the haplotype source and sink, respectively.
... The locality data, latitude and longitude were converted into XY coordinates for the analysis. Although this analysis was generally used to visualize spatial patterns of genetic diversity (Miller, 2005;Miller et al., 2006), in the several phylogeographic studies, the analysis has been applied to research past refugia (e.g., Nunome et al., 2010;Tominaga et al., 2013). Finally, we inferred the dynamics of effective population size through time in each group using the Bayesian skyline plot method (Drummond et al., 2005) as implemented in BEAST v2.6.2 (Bouckaert et al., 2019). ...
Article
Full-text available
Background: The climatic oscillations in the Quaternary period considerably shaped the distribution and population genetic structure of organisms. Studies on the historical dynamics of distribution and demography not only reflect the current geographic distribution but also allow us to understand the adaption and genetic differentiation of species. However, the process and factors affecting the present distribution and genetic structure of many taxa are still poorly understood, especially for endemic organisms to small islands. Methods: Here, we integrated population genetic and ecological niche modelling approaches to investigate the historical distribution and demographic dynamics of two co-existing salamanders on Tsushima Island, Japan: the true H. tsuensis (Group A), and Hynobius sp. (Group B). We also examined the hypothesis on the equivalency and similarity of niches of these groups by identity and background tests for ecological niche space. Results: Our result showed that Group A is considered to have undergone a recent population expansion after the Last Glacial Maximum while it is unlikely to have occurred in Group B. The highest suitability was predicted for Group A in southern Tsushima Island, whereas the northern part of Tsushima Island was the potential distribution of Group B. The results also suggested a restricted range of both salamanders during the Last Interglacial and Last Glacial Maximum, and recent expansion in Mid-Holocene. The genetic landscape-shape interpolation analysis and historical suitable area of ecological niche modelling were consistent, and suggested refugia used during glacial ages in southern part for Group A, and in northern part of Tsushima Island for Group B. Additionally, we found evidence of nonequivalence for the ecological niche of the two groups of the salamanders, although our test could not show either niche divergence or conservatism based on the background tests. The environmental predictors affecting the potential distribution of each group also showed distinctiveness, leading to differences in selecting suitable areas. Finally, the combination of population genetics and ecological modeling has revealed the differential demographic/historical response between coexisting two salamanders on a small island.
... Therefore, the JP-CB population may have experienced further geographic isolation with respect to JP-CCB and JP-YD populations, thus accumulating distinctive genetic variation (note the much longer branch of JP-CB in the phylogeny of Fig. 2B). Other studies with Japanese native species also found substantial genetic divergence between northern and central regions in Honshu, including plants (Senni et al., 2005;Ikeda et al., 2006;Hiraoka & Tomaru, 2009;Qiu et al., 2009), insects (Sota & Hayashi, 2007;Schoville et al., 2013;Saito & Tojo, 2016), and vertebrates (Oshida et al., 2009;Setiamarga et al., 2009;Nunome et al., 2010). ...
Preprint
Disjunct distribution is a key issue in biogeography and ecology, but it is often difficult to determine relative roles of dispersal vs. vicariance in disjunctions. Conandron ramondioides (Gesneriaceae) is a tertiary relict monotypic species distributed disjunctively in mainland China, Taiwan Island and Japan, where is a key region for understanding evolution and diversification of modern angiosperms. Population phylogenetic and phylogeographic structures of a comprehensive sampling of C. ramondioides by ddRAD sequencing were assessed, combined ABC modeling and SDM to infer the effects of multiple glaciation periods and to survey climatic niche differences by checking putative population divergence models and demographic scenarios. We found a very high degree of genetic differentiation among mainland China, Taiwan Island and Japan, with very limited gene flow between regions and a clear Isolation by Distance pattern. Mainland China and Japan clades diverged first from a widespread ancestral population in middle Miocene, followed by a later divergence between mainland China and Taiwan Island clades at early Pliocene. Three current groups have survived in various glacial refugia during LGM, and experienced contraction and/or bottlenecks since their divergence during Quaternary glacial cycles, with strong niche divergence between mainland China + Japan and Taiwan Island ranges. Overall, we verified a predominant role of vicariance in the current disjunction of monotypic genus Conandron. The sharp phylogenetic separation, ecological niche divergences among these three groups and the great number of private alleles in all populations sampled indicate a considerable time of independent evolution, and suggests the need of a taxonomic survey to detect potentially overlooked taxa.
... Recent northward colonization in northern Japan like that of the sub-lineage β has also been suggested for several other terrestrial vertebrates. These suggestions are often based on the "refugia" model, in which ancestral populations were confined to southern areas (refugia) during the last glacial period because of the cold climate and/or hostile vegetation in northern Japan, with rapid range expansions from these refugia to northern Japan occurring in the postglacial period (e.g., Iwasa and Abe, 2006;Kawamoto et al., 2007;Oshida et al., 2009;Nunome et al., 2010;Takehana et al., 2016). The postglacial northward expansion of the P. finitimus sub-lineage β seems concordant with this refugia scenario; however, our results also suggest persistence of the sub-lineage α in northwestern Honshu (around sites 13-20). ...
Article
We investigated the geographic diversification of Plestiodon finitimus, which occurs in the central to northern parts of the Japanese Islands, based on a time-calibrated mitochondrial DNA (mtDNA) phylogeny and external morphological characters. The mtDNA phylogeny suggests that P. finitimus diverged from its sister species Plestiodon japonicus in western Japan 2.82-4.63 million years ago (MYA), which can be explained by geographic isolation due to the spread of sedimentary basins in the Pliocene. The primary intraspecific divergence was that between P. finitimus lineages in central and northeastern Japan 1.58-2.76 MYA, which could have been caused by the upliftings of major mountain ranges. In the northeastern lineage, mtDNA and morphological characters suggest a geographic differentiation between sub-lineages of the northwestern Tohoku District (α) and other areas (β). Although the sub-lineage β occurs in a disjunct geographic range, consisting of Hokkaido and the central to south of Tohoku, these areas are bridged by populations with intermediate characteristics along the Pacific side of northern Tohoku. Overall, the geographic variation in P. finitimus in northern Japan can be explained by an initial allopatric divergence of the sub-lineages α and β at 0.71-1.39 MYA, a recent northward expansion of the sub-lineage β, and subsequent secondary introgressive hybridization between the sub-lineages.
... Pelophylax frog (Pelophylax nigromaculatus), and seed parasitic weevil (Curculio hilgendorfi) (Aoki, Kato, & Murakami, 2008;Dufresnes et al., 2016;Nagata et al., 1999;Nunome, Torii, Matsuki, Kinoshita, & Suzuki, 2010). The existence of two lineages in the aforementioned animal species was proposed based on intraspecific phylogenetic analyses, and it was tentatively explained by two biogeographi- ...
Article
Full-text available
Recent molecular studies have indicated that phylogeographical history of Japanese biota is likely shaped by geohistory along with biological events, such as distribution shifts, isolation, and divergence of populations. However, the genetic structure and phylogeographical history of terrestrial Annelida species, including leech species, are poorly understood. Therefore, we aimed to understand the genetic structure and phylogeographical history across the natural range of Haemadipsa japonica, a sanguivorous land leech species endemic to Japan, by using nine polymorphic nuclear microsatellites (nSSR) and cytochrome oxidase subunit one (COI) sequences of mitochondrial DNA (mtDNA). Analyses using nSSR revealed that H. japonica exhibited a stronger regional genetic differentiation among populations (G'ST = 0.77) than other animal species, probably because of the low mobility of land leech. Analyses using mtDNA indicated that H. japonica exhibited two distinct lineages (A and B), which were estimated to have diverged in the middle Pleistocene and probably because of range fragmentation resulting from climatic change and glacial and interglacial cycles. Lineage A was widely distributed across Japan, and lineage B was found in southwestern Japan. Analyses using nSSR revealed that lineage A was roughly divided into two population groups (i.e., northeastern and southwestern Japan); these analyses also revealed a gradual decrease in genetic diversity with increasing latitude in lineage A and a strong genetic drift in populations of northeastern Japan. Combined with the largely unresolved shallow polytomies from the mtDNA phylogeny, these results implied that lineage A may have undergone a rapid northward migration, probably during the Holocene. Then, the regional genetic structure with local unique gene pools may have been formed within each lineage because of the low mobility of this leech species.
Article
Full-text available
The Quaternary climate affected the present species richness and geographic distribution patterns of amphibians by limiting their activities during the glacial period. The present study examined the phylogenetic relationships of Japanese toads ( Bufo japonicus and B. torrenticola ) and the demography of each lineage from the past to the present based on mitochondrial sequences and ecological niche models. Japanese toads are a monophyletic group with two main clades (clades A and B). Clade A represents B. j. formosus , including three clades (clades A1, A2, and A3). Clade B contains three clades, two of which corresponded to B. j. japonicus (clades B1 and B2) and the other to B. torrenticola . Clade B2 and B. torrenticola made a sister group, and, thus, B. j. japonicus is paraphyletic. Clades A and B diverged in the late Miocene 5.7 million years ago (Mya) during the period when the Japanese archipelago was constructed. The earliest divergence between the three clades of clade A was estimated at 1.8 Mya. Clades A1 and A2 may have diverged at 0.8 Mya, resulting from the isolation in the multiple different refugia; however, the effects of the glacial climate on the divergence events of clade A3 are unclear. Divergences within clade B occurred from the late Pliocene to the early Pleistocene (3.2–2.2 Mya). Niche similarity between the parapatric clade in clade B (clades B1 and B2) indicated their allopatric divergence. It was suggested that niche segregation between B. japonicus and B. torrenticola contributed to a rapid adaptation of B. torrenticola for lotic breeding. All clade of Japanese toads retreated to each refugium at a low elevation in the glacial period, and effective population sizes increased to construct the current populations after the Last Glacial Maximum. Furthermore, we highlight the areas of climate stability from the last glacial maximum to the present that have served as the refugia of Japanese toads and, thus, affected their present distribution patterns.
Article
The Japanese serow, Capricornis crispus, is an indigenous bovid species exclusively inhabiting mountain regions in the main Japanese islands, excepting Hokkaido. It had decreased in abundance to its lowest level due to overhunting and deforestation, with its distribution severely fragmented from the middle of the 20th century, many populations of C. crispus currently facing the risk of extinction. The Kii Mountain Range (KM) on Honshu is one such location that has seen a drastic population decline of C. crispus. In this study, we examined genetic characteristics of C. crispus in KM and neighboring regions of the Chubu district, using mtDNA and microsatellite markers, in order to devise strategies for its conservation. Results for mtDNA were characterized by low nucleotide diversity with five endemic and two dominant haplotypes shared by individuals in neighboring regions. A Bayesian skyline plot indicated a gradual increase after the last glacial maximum. For microsatellites, the genetic diversity of C. crispus in KM was comparable to Shizuoka and higher than Shikoku. Recent genetic bottlenecks were strongly suggested in C. crispus in KM. Bayesian clustering showed a genetic cline between KM and neighboring regions, where multivariate analysis suggested three local populations. A Mantel test indicated male-biased dispersal. These results indicate that C. crispus in KM and neighboring regions constitute multiple local populations, connected through restricted gene flow. For the conservation of C. crispus, it is important to define small-scale conservation units, among which genetic connectivity should be facilitated to prevent further loss of genetic diversity.
Article
Largely shallow and putatively explosive divergences in the family Leporidae (rabbits and hares; order: Lagomorpha) have resulted in phylogenetic relationships that remain currently unresolved. These rapid radiations in different branches of the leporid tree have resulted in conflicting phylogenetic hypotheses. However, this phylogenetic incongruence may also result from inadequate taxon or character sampling, due to the high number of extinct and difficult to sample extant species, and highly conserved morphological characters. Sylvilagus (cottontail rabbits) constitute about 30% of the known extant leporid species. New species are routinely being recognized, and phylogenetic relationships with respect to other leporid genera, and within the genus, have failed to be recovered with certainty. Within Sylvilagus, the South American S. brasiliensis is the most widespread and poorly known taxon, likely comprising multiple species. Here, we reanalyze previously published moleculardata from phylogenetic studies on Leporidae, focusing on the S. brasiliensis group, and assessing phylogenetic relationships using bifurcating trees and split networks to identify phylogenetic regions with polytomies. We estimate differentiation and phylogenetic relationships of molecular lineages within the S. brasiliensis group. Our analyses suggest that this group contains a number of divergent taxa, well differentiated from other cottontail species. We discern two major polytomies during leporid diversification. The first, at the base of the leporid radiation, likely resulted from a combination of hard (rapid radiation) and soft polytomies (high number of unsampled extinct species). The second polytomy likely resulted from a rapid radiation during the initial diversification of the genus Sylvilagus. We conclude that only a molecular phylogeny based on a broader taxonomic representation will fully resolve leporid phylogeny.
Article
Full-text available
Patterns of geographic variation in phenotype or genotype may provide evidence for natural selection. Here, we compare phenotypic variation in color, allele frequencies of a pigmentation gene (the melanocortin-1 receptor, Mc1r), and patterns of neutral mitochondrial DNA (mtDNA) variation in rock pocket mice (Chaetodipus intermedius) across a habitat gradient in southern Arizona. Pocket mice inhabiting volcanic lava have dark coats with unbanded, uniformly melanic hairs, whereas mice from nearby light-colored granitic rocks have light coats with banded hairs. This color polymorphism is a presumed adaptation to avoid predation. Previous work has demonstrated that two Mc1r alleles, D and d, differ by four amino acids, and are responsible for the color polymorphism: DD and Dd genotypes are melanic whereas dd genotypes are light colored. To determine the frequency of the two Mc1r allelic classes across the dark-colored lava and neighboring light-colored granite, we sequenced the Mc1r gene in 175 individuals from a 35-km transect in the Pinacate lava region. We also sequenced two neutral mtDNA genes, COIII and ND3, in the same individuals. We found a strong correlation between Mc1r allele frequency and habitat color and no correlation between mtDNA markers and habitat color. Using estimates of migration from mtDNA haplotypes between dark- and light-colored sampling sites and Mc1r allele frequencies at each site, we estimated selection coefficients against mismatched Mc1r alleles, assuming a simple model of migration-selection balance. Habitat-dependent selection appears strong but asymmetric: selection is stronger against light mice on dark rock than against melanic mice on light rock. Together these results suggest that natural selection acts to match pocket mouse coat color to substrate color, despite high levels of gene flow between light and melanic populations.
Article
Full-text available
Identifying the genes underlying adaptation is a major challenge in evolutionary biology. Here, we describe the molecular changes underlying adaptive coat color variation in a natural population of rock pocket mice, Chaetodipus intermedius. Rock pocket mice are generally light-colored and live on light-colored rocks. However, populations of dark (melanic) mice are found on dark lava, and this concealing coloration provides protection from avian and mammalian predators. We conducted association studies by using markers in candidate pigmentation genes and discovered four mutations in the melanocortin-1-receptor gene, Mc1r, that seem to be responsible for adaptive melanism in one population of lava-dwelling pocket mice. Interestingly, another melanic population of these mice on a different lava flow shows no association with Mc1r mutations, indicating that adaptive dark color has evolved independently in this species through changes at different genes.
Article
Full-text available
A partial sequence of the mitochondrial cytochrome-b gene was determined for seven subspecies of sika deer (Cervus nippon) in the Japanese Islands. Nine mitochondrial DNA genotypes were distinguishable among deer sampled. Sequence analysis revealed two major phylogenetic groups comprised of northern (Hokkaido-Honshu) and southern (Kyushu) local populations. Estimated nucleotide divergence between genotypes found in southern populations was ≤1.1%, which indicated that genetic differentiation within this group occurred recently. Phylogenetic data suggest that the biogeographical boundary between northern and southern populations of sika deer lies somewhere in the Honshu mainland and not in channels that separate each Japanese Island as has been suggested. C. n. keramae, an endangered subspecies in the Kerama Islands, was found to be genotypically close to other subspecies from southern populations. C. n. keramae may have descended from deer originally introduced from Kyushu Island.
Article
We describe a model-based clustering method for using multilocus genotype data to infer population structure and assign individuals to populations. We assume a model in which there are K populations (where K may be unknown), each of which is characterized by a set of allele frequencies at each locus. Individuals in the sample are assigned (probabilistically) to populations, or jointly to two or more populations if their genotypes indicate that they are admixed. Our model does not assume a particular mutation process, and it can be applied to most of the commonly used genetic markers, provided that they are not closely linked. Applications of our method include demonstrating the presence of population structure, assigning individuals to populations, studying hybrid zones, and identifying migrants and admixed individuals. We show that the method can produce highly accurate assignments using modest numbers of loci—e.g., seven microsatellite loci in an example using genotype data from an endangered bird species. The software used for this article is available from http://www.stats.ox.ac.uk/~pritch/home.html.
Article
Most microsatellites are very polymorphic. This makes them powerful markers for observing genetic differentiation between closely related populations. The population structure of the Greenlandic Arctic fox (Alopex lagopus) was studied genetically by analysing six polymorphic microsatellite loci of 75 foxes from four populations in different parts of Greenland. Genotypes were determined at the six loci for most of the individuals. Population differentiation was quantified in three different ways both within the total population and pairwise between all populations. The tests were Fisher's exact test, Rho estimates and Fst estimates, all of which supported a highly significant subdivision of the total population, and they showed significant differentiation in allele frequencies between all pairs of localities. It is concluded that the known long-distance migration of the Greenlandic Arctic fox has not resulted in complete genetic mixing of the populations. Fisher's exact test was also used to estimate levels of genetic differentiation between the two colour morphs: white and blue. No difference was found between allele frequencies of the two color morphs in any of the locations, and it was concluded that the white and blue morphs of the Greenlandic Arctic fox share the same habitat, at least during the mating season.
Article
Tohoku hare, like other wild mammals, manifests periodical sexual activity according to the season. Minimal ovarian weight is observed during the season from November until January. Generally, gradual enlargement of size with increase in weight is seen from the end of Feburuary or from early March. During the breeding season, in spring and summer, the ovary is large in size and active. It undergoes rapid degeneration from September or sometimes from October. In the present experiment the effect of prolonged light exposure upon breeding was studied, with special reference to ovarian activity. From December 11, 1968, 25 mature females and 10 males in captivity were exposed daily to natural daylight from sunrise to 9a.m. and then artificial light was added until 9p.m. An increase in ovarian weight was observed from 50 days after the start of the experiment, namely from the end of January. However, a similar state of ovarian activity is not normally attained until late Feburuary or early March in the wild. Though the exposure experiment was ended on May 6, the ovary showed no signs of degeneration even on Apil 12. This year, the first pregnant female was captured in the wild on Feburuary 26 and one female gave birth to a litter of 3 on April 5. However, in the experiment a pregnant female was observed on January 30 and one female gave birth to a litter of one on Feburuary 22. The experimented animals continued to give birth until April 10. From April 11 to June 29 breeding seemed to be ceased but from June 30 it began again and continued until August 22. It is worthy to note that the ovaries, both of the wild animals or of the animals experimented, contained fully grown follicles at all seasons of the year. © 1971, JAPANESE SOCIETY OF APPLIED ENTOMOLOGY AND ZOOLOGY. All rights reserved.
Article
A new method called the neighbor-joining method is proposed for reconstructing phylogenetic trees from evolutionary distance data. The principle of this method is to find pairs of operational taxonomic units (OTUs [= neighbors]) that minimize the total branch length at each stage of clustering of OTUs starting with a starlike tree. The branch lengths as well as the topology of a parsimonious tree can quickly be obtained by using this method. Using computer simulation, we studied the efficiency of this method in obtaining the correct unrooted tree in comparison with that of five other tree-making methods: the unweighted pair group method of analysis, Farris's method, Sattath and Tversky's method, Li's method, and Tateno et al.'s modified Farris method. The new, neighbor-joining method and Sattath and Tversky's method are shown to be generally better than the other methods.