large‐scale migrations of brown bears in eurasia...

12
ORIGINAL ARTICLE Large-scale migrations of brown bears in Eurasia and to North America during the Late Pleistocene Peeter Anijalg 1 | Simon Y. W. Ho 2 | John Davison 1 | Marju Keis 1 | Egle Tammeleht 1 | Katalina Bobowik 2 | Igor L. Tumanov 3 | Alexander P. Saveljev 4 | Elena A. Lyapunova 5 | Alexandr A. Vorobiev 6 | Nikolai I. Markov 6 | Alexey P. Kryukov 7 | Ilpo Kojola 8 | Jon E. Swenson 9,10 | Snorre B. Hagen 11 | Hans Geir Eiken 11 | Ladislav Paule 12 | Urmas Saarma 1 1 Department of Zoology, Institute of Ecology and Earth Sciences, University of Tartu, Tartu, Estonia 2 School of Life and Environmental Sciences, University of Sydney, Sydney, NSW, Australia 3 West Branch, Russian Research Institute of Game Management and Fur Farming, Saint Petersburg, Russia 4 Russian Research Institute of Game Management and Fur Farming, Kirov, Russia 5 N.K. Koltzov Institute of Developmental Biology, Russian Academy of Sciences, Moscow, Russia 6 Institute of Plant and Animal Ecology, Ural Branch of Russian Academy of Sciences, Yekaterinburg, Russia 7 Federal Scientific Center of the East Asia Terrestrial Biodiversity, Far Eastern Branch of the Russian Academy of Sciences, Vladivostok, Russia 8 Natural Resources Institute, Rovaniemi, Finland 9 Faculty of Environmental Sciences and Natural Resource Management, Norwegian University of Life Sciences, As, Norway 10 Norwegian Institute for Nature Research, Trondheim, Norway 11 NIBIO, Norwegian Institute of Bioeconomy Research, Svanvik, Norway 12 Faculty of Forestry, Technical University, Zvolen, Slovakia Correspondence Urmas Saarma, Department of Zoology, Institute of Ecology and Earth Sciences, University of Tartu, Tartu, Estonia. Email: [email protected] Funding information Seventh Framework Programme, Grant/ Award Number: contracts 247652 and 2096/7. PR UE/2011/2; Estonian Ministry of Education and Research, Grant/Award Number: IUT20-32 Editor: Gerald Schneeweiss Abstract Aim: Climatic changes during the Late Pleistocene had major impacts on popula- tions of plant and animal species. Brown bears and other large mammals are likely to have experienced analogous ecological pressures and phylogeographical pro- cesses. Here, we address several unresolved issues regarding the Late Pleistocene demography of brown bears: (1) the putative locations of refugia; (2) the direction of migrations across Eurasia and into North America; and (3) parallels with the demographic histories of other wild mammals and modern humans. Location: Eurasia and North America. Methods: We sequenced 110 complete mitochondrial genomes from Eurasian brown bears and combined these with published sequences from 138 brown bears and 33 polar bears. We used a Bayesian approach to obtain a joint estimate of the phylogeny and evolutionary divergence times. The inferred mutation rate was com- pared with estimates obtained using two additional methods. Results: Bayesian phylogenetic analysis identified seven clades of brown bears, with most individuals belonging to a very large Holarctic clade. Bears from the wide- spread clade 3a1, which has a distribution from Europe across Asia to Alaska, shared a common ancestor about 45,000 years ago. DOI: 10.1111/jbi.13126 394 | © 2017 John Wiley & Sons Ltd wileyonlinelibrary.com/journal/jbi Journal of Biogeography. 2018;45:394405.

Upload: others

Post on 28-Jun-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

OR I G I N A L A R T I C L E

Large-scale migrations of brown bears in Eurasia and to NorthAmerica during the Late Pleistocene

Peeter Anijalg1 | Simon Y. W. Ho2 | John Davison1 | Marju Keis1 | Egle Tammeleht1 |

Katalina Bobowik2 | Igor L. Tumanov3 | Alexander P. Saveljev4 | Elena A. Lyapunova5 |

Alexandr A. Vorobiev6 | Nikolai I. Markov6 | Alexey P. Kryukov7 | Ilpo Kojola8 |

Jon E. Swenson9,10 | Snorre B. Hagen11 | Hans Geir Eiken11 | Ladislav Paule12 |

Urmas Saarma1

1Department of Zoology, Institute of Ecology and Earth Sciences, University of Tartu, Tartu, Estonia

2School of Life and Environmental Sciences, University of Sydney, Sydney, NSW, Australia

3West Branch, Russian Research Institute of Game Management and Fur Farming, Saint Petersburg, Russia

4Russian Research Institute of Game Management and Fur Farming, Kirov, Russia

5N.K. Koltzov Institute of Developmental Biology, Russian Academy of Sciences, Moscow, Russia

6Institute of Plant and Animal Ecology, Ural Branch of Russian Academy of Sciences, Yekaterinburg, Russia

7Federal Scientific Center of the East Asia Terrestrial Biodiversity, Far Eastern Branch of the Russian Academy of Sciences, Vladivostok, Russia

8Natural Resources Institute, Rovaniemi, Finland

9Faculty of Environmental Sciences and Natural Resource Management, Norwegian University of Life Sciences, �As, Norway

10Norwegian Institute for Nature Research, Trondheim, Norway

11NIBIO, Norwegian Institute of Bioeconomy Research, Svanvik, Norway

12Faculty of Forestry, Technical University, Zvolen, Slovakia

Correspondence

Urmas Saarma, Department of Zoology,

Institute of Ecology and Earth Sciences,

University of Tartu, Tartu, Estonia.

Email: [email protected]

Funding information

Seventh Framework Programme, Grant/

Award Number: contracts 247652 and

2096/7. PR UE/2011/2; Estonian Ministry

of Education and Research, Grant/Award

Number: IUT20-32

Editor: Gerald Schneeweiss

Abstract

Aim: Climatic changes during the Late Pleistocene had major impacts on popula-

tions of plant and animal species. Brown bears and other large mammals are likely

to have experienced analogous ecological pressures and phylogeographical pro-

cesses. Here, we address several unresolved issues regarding the Late Pleistocene

demography of brown bears: (1) the putative locations of refugia; (2) the direction

of migrations across Eurasia and into North America; and (3) parallels with the

demographic histories of other wild mammals and modern humans.

Location: Eurasia and North America.

Methods: We sequenced 110 complete mitochondrial genomes from Eurasian

brown bears and combined these with published sequences from 138 brown bears

and 33 polar bears. We used a Bayesian approach to obtain a joint estimate of the

phylogeny and evolutionary divergence times. The inferred mutation rate was com-

pared with estimates obtained using two additional methods.

Results: Bayesian phylogenetic analysis identified seven clades of brown bears, with

most individuals belonging to a very large Holarctic clade. Bears from the wide-

spread clade 3a1, which has a distribution from Europe across Asia to Alaska, shared

a common ancestor about 45,000 years ago.

DOI: 10.1111/jbi.13126

394 | © 2017 John Wiley & Sons Ltd wileyonlinelibrary.com/journal/jbi Journal of Biogeography. 2018;45:394–405.

Page 2: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

Main conclusions: We suggest that the Altai-Sayan region and Beringia were impor-

tant Late Pleistocene refuge areas for brown bears and propose large-scale migra-

tion scenarios for bears in Eurasia and to North America. We also argue that brown

bears and modern humans experienced a demographic standstill in Beringia before

colonizing North America.

K E YWORD S

Altai, Beringia, climate change, mitochondrial genome, molecular clock, phylogeography, Ursus

arctos

1 | INTRODUCTION

Climatic fluctuations during the Pleistocene have shaped the distri-

bution, genetic structure, and diversity of present-day species in Eur-

asia and North America. During the Last Glacial Maximum (LGM;

26.5–19 kya), many high-latitude areas were covered by continental

ice sheets and temperate species were largely restricted to warmer

and more hospitable areas known as glacial refugia (Hofreiter & Ste-

wart, 2009). Some southern parts of Europe, such as the Mediter-

ranean peninsulas, have long been recognized as LGM refugia for

temperate species. There is also growing recognition of more north-

erly refugia in Europe, including the Carpathian Mountains (Def-

fontaine et al., 2005; Kotl�ık et al., 2006; Saarma et al., 2007; Schmitt

& Varga, 2012).

In phylogeographical studies of large mammals, the brown bear

(Ursus arctos L.) has been a model species. This is largely due to its

wide past and present distribution, as well as female philopatry (Davi-

son et al., 2011). In historical times, brown bears were common in

most parts of Eurasia and in the western part of North America down

to northern Mexico (Kurt�en, 1968). Present-day bear populations

occupy somewhat smaller areas, with many populations fragmented

mainly because of human persecution and loss of suitable habitat

(Davison et al., 2011; Swenson, Taberlet, & Bellemain, 2011). The lar-

gest brown bear population in northern Eurasia stretches from Scan-

dinavia to the Russian Far East and is associated predominantly with

the occurrence of boreal and temperate forests. The range is

restricted by tundra in the north and by dry steppe in the south

(Servheen, Herrero, & Peyton, 1999). However, the Gobi bear in

Mongolia is desert-dwelling (McCarthy, Waits, & Mijiddorj, 2009),

indicating that brown bears can survive in fairly hostile environments.

For many species, including the brown bear, long-term isolation

of populations in separated glacial refugia resulted in genetic diver-

gence among populations, loss of genetic variability and complex

patterns of migrations during the Late Pleistocene and Holocene

(Davison et al., 2011; Hewitt, 2000; Taberlet, Fumagalli, Wust-Saucy,

& Cosson, 1998). However, despite the wide Eurasian distributions

of many species, including several model species for biogeography,

we lack an equivalent understanding of the processes influencing

populations in Asia.

Analysis of genetic markers has provided a highly informative

means of reconstructing both ancient and contemporary patterns of

diversity and demography. Important insights into geographical varia-

tion among populations of animals, such as those of the brown bear,

have been gained through analyses of autosomal loci (Swenson

et al., 2011; Tammeleht et al., 2010), Y-chromosome markers (Bidon

et al., 2014; Schregel et al., 2015), and other sources of data, such

as diet (Vulla et al., 2009). Based on complete mitogenome

sequences of brown bears in north-western Eurasia, Keis et al.

(2013) identified population structuring, demographic history, and

spatial population processes that had not previously been detected

using shorter sequences.

We expect that a wider analysis of mitogenomes from Eurasia to

North America will help to expose new details of ancient population

processes and their time-frame. Mitochondrial DNA has been a pre-

ferred genetic marker for studying bears for several reasons. Perhaps

the most relevant to phylogeographical studies is that because

female brown bears are philopatric, the phylogeographical signal is

better preserved in mitochondrial DNA than in autosomal markers or

the Y-chromosome. However, the phylogeographic history of the

female lineage does not necessarily coincide with that of the whole

species (Davison et al., 2011). A notable example of this is the mito-

chondrial paraphyly of brown bears (Talbot & Shields, 1996), which

stands in contrast with the clear support for monophyly from male-

specific (Bidon et al., 2014) and biparental markers (Liu et al., 2014).

The present-day European population of brown bears consists of

two main mitochondrial lineages, the eastern (clade 3a) and western

(clade 1), of which the latter is divided into subclades 1a and 1b

(Taberlet & Bouvet, 1994). These clades and subclades exhibit con-

siderable geographical differentiation. Subclade 1a originated from

an Iberian refuge and is now found in southwestern Europe and

southern Scandinavia; however, the recolonization of southern Scan-

dinavia was most likely not limited to migrations from a single refu-

gium (Bray et al., 2013). In contrast, subclade 1b has virtually

remained in its refugial territory in the south and east of Europe

(Italy and Balkans). In North America, four mitochondrial clades have

been identified in contemporary brown bears: clade 2a on the Admi-

ralty, Baranof, and Chichagof (ABC) islands of Alaska, clade 3a1 in

western Alaska, clade 3b in eastern Alaska, and clade 4 in the south-

ern distribution of the species in continental North America (Davison

et al., 2011; Hirata et al., 2013).

The eastern clade present in Europe (3a) divides further into sub-

clades 3a1 and 3a2 (Hirata et al., 2013). Subclade 3a1 extends from

ANIJALG ET AL. | 395

Page 3: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

Fennoscandia to East Asia, and into North America (Keis et al.,

2013; Korsten et al., 2009; Murtskhvaladze, Gavashelishvili, & Tarkh-

nishvili, 2010; Saarma & Kojola, 2007; Waits, Talbot, Ward, &

Shields, 1998). Other clades in Asia appear geographically rather

confined: clade 3a2 to central Hokkaido, 3b to eastern Hokkaido,

the Russian Far East and southern Siberia (Tomsk region), clade 4 to

southern Hokkaido, and clade 5 to Tibet (Guskov, Sheremeteva,

Seredkin, & Kryukov, 2013; Hirata et al., 2013; Salomashkina, Kholo-

dova, Tutenkov, Moskvitina, & Erokhin, 2014).

Although broad patterns of mitogenomic composition are known

for Asian brown bear populations, phylogeographical syntheses of

these patterns remain tentative. The widespread sublineage 3a1 is

believed to have emerged from multiple glacial refugia, but the loca-

tions of putative areas in Asia remain unknown. The Siberian plains

and mountains were largely unglaciated during the LGM, represent-

ing potential refugia (Davison et al., 2011). Indeed, several studies

have suggested that the Altai-Sayan region in south-central Siberia

was one of the most important Asian refugia during the Late Pleis-

tocene, due to its stable climatic conditions and species richness

throughout this period of time (Pavelkov�a �Ri�c�ankov�a, Robovsk�y, &

Riegert, 2014; Pavelkov�a �Ri�c�ankov�a, Robovsk�y, Riegert, & Zrzav�y,

2015). Moreover, modern humans are known to have inhabited the

Altai region for at least 45 kyr (Goebel, 1999) and the area played a

significant role in the peopling of Siberia and North America (Dulik

et al., 2012). Reconstructing the migrations of modern humans dur-

ing the Late Pleistocene and early Holocene (c. 50–10 kya) poses a

major challenge, because of the small population size and low den-

sity of archaeological sites during this period. However, other large

mammal species that shared similar ecological demands as modern

humans, such as the brown bear, wapiti (Cervus canadensis; Meiri

et al., 2014) and grey wolf (Canis lupus; Koblm€uller et al., 2016), rep-

resent a largely unexploited source of information that can also

improve our understanding of the drivers of ancient population pro-

cesses in modern humans.

Several other geographical regions are of potential relevance for

understanding genetic diversity in Asia and links with the adjacent

regions of Europe and North America. First, there is evidence to sug-

gest that the Ural Mountains acted as a glacial refugium for animals

and modern humans at the border of Europe and Asia (Svendsen

et al., 2010). Another refugium is believed to have been located in

Beringia, the area stretching from the Verkhoyansk Mountains and

Lena River Basin in eastern Eurasia to Alaska and the Northwest

Territories of Canada in North America. Compared with areas of sim-

ilar latitude, the climate in Beringia stayed relatively mild and ice-free

during the LGM due to the warm North Pacific circulation (Lambeck,

Yokoyama, & Purcell, 2002). The Bering Land Bridge, which emerged

during times of low sea level, allowed migrations of mammals

between Eurasia and North America several times during the Pleis-

tocene.

The aim of this study was to investigate the Late Pleistocene

phylogeographical history of brown bears in Eurasia and their colo-

nization of North America. We sequenced the complete mitochon-

drial genomes of 110 brown bears across Eurasia and included

publicly available data from Eurasia to North America in a dated phy-

logenetic analysis. We hypothesized that: (1) the widely distributed

clade 3a1 would display Eurasia-wide structure that has not been

apparent using lower-resolution genetic markers; and (2) populations

were restricted to refugia in the areas of Altai, Beringia and the

Urals, such that the ages of the clades containing individuals from

these areas are consistent with the timing of glacial restriction. We

also compared our data with existing information from modern

humans and other mammal species to assess the generality of Late

Pleistocene migration scenarios.

2 | MATERIALS AND METHODS

Tissue samples from 110 brown bears, legally harvested between

1999 and 2010 for purposes other than this study, were used for

full mitochondrial DNA sequencing. These included three from

Romania, five from Estonia, five from Sweden, 10 from Finland and

87 from the Russian Federation (Figure 1). For DNA extraction, we

used the High PCR Template Preparation Kit (Roche Diagnostics,

Mannheim, Germany) following the manufacturer’s instructions.

PCR and sequencing of the complete mitochondrial genomes

was done as described by Keis et al. (2013), with a modified set of

primers; we replaced six primers with newly designed ones to

improve mitogenome assembly (Table S1 in Appendix S1). All

sequences have been deposited in GenBank (accession numbers

KY419593–KY419702). For phylogenetic analysis, we added

sequences from 138 brown bears and 33 polar bears (U. maritimus)

from GenBank (Table S4 in Appendix S1). In total, our data set con-

tained 281 mitogenome sequences from brown bears and polar

bears.

To estimate the matrilineal phylogeny and evolutionary time-

scale of brown bears, we analysed complete mitogenomes using a

Bayesian approach in BEAST 1.8.4 (Drummond, Suchard, Xie, & Ram-

baut, 2012). The best-fitting partitioning scheme was chosen based

on the Bayesian information criterion using PartitionFinder (Lanfear,

Calcott, Ho, & Guindon, 2012), which partitioned the alignment

into six subsets: D-loop, rRNA genes, tRNA genes, and the three

codon positions of all 13 protein-coding genes. We compared con-

stant-size and skyride coalescent models for the tree prior using

Bayes factors in BEAST, calculated using the stepping-stone estimator

of the marginal likelihood (Xie, Lewis, Fan, Kuo, & Chen, 2011).

The constant-size model was decisively favoured (BF > 100) and

we used this to provide the prior for the tree topology and relative

node times.

We analysed the partitioned alignment using a strict molecular

clock, owing to the intraspecific nature of the data. Preliminary anal-

yses using a more parameter-rich uncorrelated lognormal relaxed

clock led to poor mixing and convergence in our analyses. A rate

multiplier was specified for each data subset. The molecular clock

was calibrated by the inclusion of a sequence from an ancient polar

bear (age 120,000 years; GU573488). Estimates of posterior distri-

butions were obtained via Markov chain Monte Carlo sampling, with

396 | ANIJALG ET AL.

Page 4: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

samples drawn every 5,000 steps over a total of 50,000,000 steps.

Each analysis was run in duplicate and the samples were combined

after removing the first 10% as burn-in. We checked that the effec-

tive sample sizes of all parameters were >200.

We performed a date-randomization test to investigate the

performance of the sequence ages as calibrations for the molecu-

lar clock (Ramsden, Holmes, & Charleston, 2009) and generated

10 replicate data sets in which the sequence ages were randomly

reassigned. The mean rate estimate from the original data set was

not included in the 95% credibility intervals (CI) of the rate esti-

mates from any of the 10 date-randomized replicates. Further-

more, the rate estimates from the date-randomized replicates had

long lower tails that spanned multiple orders of magnitude. In

combination, these features suggest that the original data set has

sufficient temporal structure and spread for calibrating the esti-

mate of the rate (Duchene, Duchene, Holmes, & Ho, 2015; Ho

et al., 2011).

As an independent check on the reliability of the rate estimate,

we analysed the data using two other tip-dating methods: root-to-

tip regression and least-squares dating. For both of these methods,

we inferred a phylogenetic tree using maximum likelihood with 100

random starts in RAxML 8.2.4 (Stamatakis, 2014). The inferred phy-

logram was then used for regression of root-to-tip distances against

sampling time in TempEst (Rambaut, Lam, Carvalho, & Pybus, 2016)

and for least-squares dating in LSD (To, Jung, Lycett, & Gascuel,

2016).

3 | RESULTS

Our Bayesian phylogenetic analysis of 281 mitogenome sequences

from brown bears to polar bears distinguished eight clades (Fig-

ure 2). According to the prevailing terminology (Davison et al., 2011;

Hirata et al., 2013), these clades are: 1 and 2a (and 2b for polar

bears) representing the western lineage; and 3a1, 3a2, 3b, 4 and 5

representing the eastern lineage. All newly sequenced samples fell

into the widely distributed clade 3a1, except for one sample from

Sweden which belonged to western clade 1.

In our Bayesian phylogenetic analysis, the mitogenomic mutation

rate was estimated at 2.48 9 10�8 mutations per site per year, with a

95% CI of 1.93 9 10�8 to 3.04 9 10�8 mutations per site per year.

This estimate was slightly lower than the estimate based on regression

of root-to-tip distances against sample age (3.35 9 10�8 mutations

per site per year), but similar to the rate estimated using least-squares

dating (2.84 9 10�8 mutations per site per year).

The most recent common ancestor (MRCA) of clade 3a1 lived

c. 45 kya (95% CI: 31–58 kya) (Figure 2; see also Table S2 and

Table S3 for 95% CIs, in Appendix S1). This split separated brown

bears currently inhabiting Romania and Bulgaria from the rest of the

3a1 clade. The latter diverged 37 kya (95% CI: 27–48 kya) into two

large subclades: (1) a Kamchatkan–Alaskan subclade comprising bears

from Alaska and Kamchatka (34 kya, 95% CI: 24–44 kya); and (2) a

Eurasian subclade that includes bears inhabiting the vast area from

eastern Europe to the Russian Far East (32 kya, 95% CI 23–42 kya).

F IGURE 1 Map showing locations ofEurasian and North American brown bearsanalysed in this study: 110 newlysequenced mitogenomes (circles) and 138from databases (triangles). Colourscorrespond to clades in Figure 2[Colour figure can be viewed atwileyonlinelibrary.com]

ANIJALG ET AL. | 397

Page 5: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

Following the formation of the Kamchatkan-Alaskan subclade,

the Alaskan bears separated from the Kamchatkan bears (27 kya,

95% CI 19–36 kya). The latter formed two clusters: “to E-Kam-

chatka” (14 kya, 95% CI 7–21 kya), consisting mostly of bears

inhabiting the north-eastern part of the peninsula; and “to W-Kam-

chatka” (25 kya, 95% CI 17–33 kya), largely represented by bears

from the south-western part of the peninsula. The Eurasian subclade

divided into lineages that reached Primorye, Sakhalin, and Kam-

chatka (27 kya, 95% CI 18–37 kya) and one that spread widely in

Eurasia (29 kya, 95% CI 21–38 kya). The latter divided further into

two: bears that eventually reached Europe and bears that migrated

to the Urals and to Magadan region, with only minimal overlap in

the Perm and Sverdlovsk regions.

4 | DISCUSSION

Although the phylogeography of European mammals has been stud-

ied fairly extensively, we have only a fragmentary understanding of

the phylogeographical processes affecting mammals in the vast terri-

tory of northern Eurasia. An important step towards improving this

knowledge is to gain a reliable estimate of the evolutionary rate and

time-scale of each species (Ho, Saarma, Barnett, Haile, & Shapiro,

2008). Our analyses confirm previous suggestions that a single

ancient DNA sequence can be sufficient for calibrating the molecular

clock, provided that it is sufficiently old in comparison with the evo-

lutionary time-scale of the samples in the data set (Molak, Lorenzen,

Shapiro, & Ho, 2013). This is further supported by the consistency in

our estimates of the mutation rate using three different methods.

To date, only shallow regional genetic structure has been

detected among several mammal species occurring across northern

Eurasia (e.g. Hindrikson et al., 2017; Korsten et al., 2009). However,

that emerging picture is by no means definitive, as these compar-

isons were made using relatively short mitochondrial sequences

(Davison et al., 2011). Based on our analyses of complete mitogen-

ome data, we propose several scenarios of brown bear migration

across Eurasia and to North America during the last 50 kyr

(Figure 3).

F IGURE 2 Bayesian phylogenetic tree of brown bears inferred from mitochondrial genomes. The numbers on nodes and after cladesrepresent posterior mean estimates of the ages of different bear lineages, with numbers in parentheses representing the 95% credibilityintervals. Red circles at nodes represent posterior probabilities >0.95. The blue vertical columns represent long cold periods according to Ahnand Brook (2014). The tree is drawn to a time-scale, with node ages given in thousands of years before present. Colours indicate differentmitochondrial clades. Colours of clades correspond to those in Figure 1 [Colour figure can be viewed at wileyonlinelibrary.com]

398 | ANIJALG ET AL.

Page 6: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

4.1 | Spatial distribution of bear clade 3a in theLate Pleistocene

The full distribution of clade 3a bears around 50 kya remains

unknown, but existing data indicate that the distribution could have

been wide. Two radiocarbon-dated 3a bear samples are known from

this period, both c. 47 ky old: one from Ramesch Cave, Austria

(Hofreiter et al., 2004) and the other from Tain Cave in the Urals,

Russia (Bray, 2010). The distribution of clade 3a probably reached

even farther east. One of the arguments to support this is the

F IGURE 3 Possible distribution and migration scenarios of brown bears from clade 3a1 in the Late Pleistocene. Colours correspond tothose in Figures 1 and 2. Excerpts of the corresponding parts of the phylogenetic tree (according to Figure 2) are shown on the right of eachpanel. Subfossils are marked by large brown circles [Colour figure can be viewed at wileyonlinelibrary.com]

ANIJALG ET AL. | 399

Page 7: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

divergence of 3a1 and 3a2 that started c. 53 kya (95% CI: 38–

71 kya). As 3a2 bears are found only in Hokkaido (Hirata et al.,

2013), the likely distribution of 3a bears before the split extended to

eastern Asia. The vast distribution is also supported by the biotope

map in Hoogakker et al. (2016) which shows that around 54 kya

boreal and temperate forests covered a continuous area from west-

ern Europe to the Russian Far East (Figure 4c). The wide distribution

of the 3a lineage might have been ruptured by significant global cli-

mate cooling about 48–54 kya (Kindler et al., 2014), with these

bears potentially restricted to different refugia (Figure 3a).

The most recent common ancestor of clade 3a1 lived 45 kya

(95% CI: 31–58 kya). This clade split into two mitochondrial lineages,

the first of which is represented by bears from Romania to Bulgaria.

These might be descendants of ancient bears that lived in central

Europe c. 50 kya (e.g. the individual from Ramesch Cave) that found

refuge south of the Carpathian Mountains. The Carpathian region

probably acted as a glacial refugium for brown bears, as previously

suggested by Sommer and Benecke (2005) and Saarma et al. (2007),

and also for other species, such as the common vole (Microtus arva-

lis) (Stojak, McDevitt, Herman, Searle, & W�ojcik, 2015) and weasel

(Mustela nivalis) (McDevitt et al., 2012). Two other Romanian sam-

ples belonged to the much larger second lineage of 3a1 (Figure 2),

which moved across Eurasia and to North America. This suggests

that bears made their way to the Carpathians (Romania) not only

from the north (central Europe), but also from the east (from the

Altai-Sayan region, discussed below).

The clade that spread throughout Eurasia and part of North

America shared a most recent common ancestor about 37 kya (95%

CI: 27–48 kya; Figure 3b). Several lines of evidence suggest that this

ancestral population was restricted to a refuge area in the mountain-

ous Altai-Sayan region of Central Asia. This region is known as a bio-

diversity hotspot for temperate flora and fauna, having maintained

high ecological stability since the Late Pleistocene (Allen et al., 2010;

Huntley et al., 2013). Detailed analysis of the Late Pleistocene mam-

mal assemblages in the Altai region has shown that no significant

changes in composition or ecological structure occurred between the

F IGURE 4 Maps showing the extent ofdifferent biotopes (a) 6 kya, (b) 21 kya and(c) 54 kya (modified from Hoogakker et al.,2016). Note that the precise distribution ofbiotypes – for example the extent of forestin central Eurasia in periods (b) and (c) –comes with a degree of uncertainty (seeHoogakker et al., 2016 for discussion)Potential refuge areas are shown withdashed lines [Colour figure can be viewedat wileyonlinelibrary.com]

400 | ANIJALG ET AL.

Page 8: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

Late Pleistocene and the Holocene, providing a modern analogue for

the Pleistocene communities (Agadjanian & Serdyuk, 2005;

Pavelkov�a �Ri�c�ankov�a et al., 2014, 2015).

The Altai-Sayan region is at the geographical centre of the brown

bear 3a1 distribution, making it the most parsimonious location for

the ancestral refuge area. Furthermore, historical bear samples exhi-

bit higher genetic diversity in this region than elsewhere in northern

Eurasia (Hirata, Abramov, Baryshnikov, & Masuda, 2014). The Altai-

Sayan region is likely to have served as an important glacial refugium

for modern humans and as an important route for the peopling of

northern Asia (Derenko et al., 2014).

4.2 | Colonizing North America

One of the major subclades of 3a1 bears headed towards Beringia,

eventually colonizing Kamchatka and Alaska. Bears from the other

major subclade of 3a1 dispersed widely throughout Eurasia. Interest-

ingly, bears in the Magadan area, a region close to Kamchatka, were

neither part of the migration wave that colonized the Kamchatka

Peninsula, nor part of the wave that entered North America. Rather,

these bears apparently did not move farther north-east. However,

we found a bear from the east Asian migration in Kamchatka (Fig-

ure 3c), suggesting that some bears from this population eventually

reached Kamchatka, perhaps somewhat later than bears that were

part of the earlier migration that colonized North America.

Brown bears belonging to the Kamchatkan-Alaskan mitochondrial

subclade apparently entered western Beringia before the beginning

of the LGM, because north-eastern territories outside Beringia were

mainly polar desert during the LGM and unsuitable for brown bears

(Ray & Adams, 2001). Indeed, Pitulko et al. (2004) described c. 34-

kyr-old brown bear remains at the Yana site (Yana RHS), whereas

Bray (2010) was able to sequence a 135-bp mitochondrial fragment

of a c. 30-kyr-old brown bear sample from Indigirka area that

belonged to clade 3a. Both sites are in western Beringia, providing

evidence that brown bears were present there before the LGM.

Moreover, our data suggest that the Kamchatkan-Alaskan subclade

also reached western Beringia around that time (see discussion

about Kamchatka below). There is no biotope map relating to this

time period, but the 54 kya map shows an almost continuous distri-

bution of boreal forest from Eurasia to central Beringia (Figure 4c).

Around 37 kya this distribution was probably interrupted by abrupt

climate change, thus separating the Kamchatkan-Alaskan subclade.

The MRCA of the Alaskan brown bears lived 22 kya (95% CI:

13–32 kya; Figure 2). The most likely location of these bears was

west of the current Bering Strait, where they remained for about 6–

7 thousand years (Beringian Standstill), before crossing the Bering

Land Bridge and reaching Alaska c. 15 kya. The Beringian Standstill

hypothesis was originally proposed for migrations of modern

humans, whereby the ancestors of Native Americans remained in

western Beringia for thousands of years before migrating to North

America (Tamm et al., 2007). During this time the climatic conditions

were unsuitable for modern humans at higher latitudes outside the

land bridge (Hoffecker, Elias, & O’Rourke, 2014). The biotope map

describing the LGM (Figure 4b) shows that central Beringia was cov-

ered by boreal forest, providing suitable habitat for brown bears.

There has been growing genetic support for the standstill hypothesis

(Watson, 2017), including evidence from ancient mitochondrial gen-

omes from modern humans (Llamas et al., 2016). A study of wapiti

also found that this species was present in western Beringia

c. 50 kya, but did not advance to North America until 15 kya (Meiri

et al., 2014).

The favourable conditions for many plant and animal species in

western and central Beringia during the LGM are believed to have

been caused by the North Pacific circulation, which brought warm

and moist air to the area (Brubaker, Anderson, Edwards, & Lozhkin,

2005; Hoffecker et al., 2014; Willerslev et al., 2014). However, such

good conditions were not available in eastern Beringia before

15 kya. Here the environment was apparently a cold, semi-arid, and

treeless steppe-tundra, suitable only for species adapted to extreme

cold, such as woolly mammoths (Mammuthus primigenius) (Guthrie,

2001).

Synchrony in the timing of colonization and expansion implies

that the ecology of large mammals, such as the wapiti and brown

bear, includes key limiting factors that can enhance our understand-

ing of the ancient movements of modern humans (Meiri et al.,

2014). However, despite similarities between the migrations of

brown bears and modern humans towards North America c. 15 kya,

there are also some important differences. Brown bears are known

to have colonized North America earlier; clades 2c, 3c, and 4 were

already in North America before the last glaciation (Barnes, Matheus,

Shapiro, Jensen, & Cooper, 2002; Davison et al., 2011). The genetic

diversity of Beringian brown bears in the pre-glacial period

(c. 40 kya) was higher than that of modern North American popula-

tions, but by 15 kya the genetic diversity was similar to that of con-

temporary bears in the New World (Leonard, Wayne, & Cooper,

2000).

4.3 | Colonizing the Kamchatka Peninsula

The extent and timing of the last glaciation in Kamchatka were lar-

gely influenced by the Laurentide Ice Sheet (Barr & Solomina, 2014).

The ice is believed to have advanced in two major stages in Late

Pleistocene Kamchatka: the first advance was extensive and

occurred between 60 and 31 kya; the more recent advance was less

extensive, characterized primarily by mountain glaciation, and coin-

cided with the LGM. The MRCA of brown bears that colonized

North America and Kamchatka was dated to 34 kya (95% CI: 24–

44 kya), coinciding with the end of the major glaciation on Kam-

chatka. At this point, bears would have been able to occupy the ter-

ritory in the vicinity of the peninsula, or even enter the peninsula to

some extent. The biotope map (Figure 4b) shows that boreal forests

were present to a limited extent in northern Kamchatka during the

LGM, making it a suitable habitat for brown bears that later colo-

nized the peninsula.

Kurt�en (1968) mentioned that the American broad-skulled bears

morphologically resemble the Kamchatkan bears, whereas the

ANIJALG ET AL. | 401

Page 9: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

American narrow-skulled bears are more similar to the bears from

north-eastern Siberia. The morphological resemblance might be the

result of diet, because the largest bears occur in regions where they

have access to large amounts of meat, especially salmon (Hilderbrand

et al., 1999). Another mammal species, the northern red-backed vole

(Myodes rutilus), exhibits a similar Kamchatkan-Alaskan clade (Kohli,

Fedorov, Waltari, & Cook, 2015). The close phylogenetic association

between Alaskan and Kamchatkan populations in these two species

points to a migration pattern that might also be shared by other

mammals.

4.4 | Large-scale migrations across Siberia andexpansion to Europe

The MRCA of the Eurasian subclades was dated to 32 kya (95% CI:

23–42 kya). At this point, most of the Eurasian bears split from the

branch that today can be found in the Primorye region, Sakhalin

Island and Kamchatka, suggesting that at least some bears managed

to reach Kamchatka after the initial colonization of the peninsula

(Figure 3c). Bears from this subclade were probably restricted to a

refuge area in the Far East during the LGM, a plausible region being

the Manchurian Plain. According to the climate model by Hoogakker

et al. (2016), around 21 kya the Manchurian Plain was suitable for

boreal and temperate forests, making it potentially habitable for

brown bears (Figure 4b).

The other Eurasian branch split into two around 29 kya (95% CI:

21–38 kya; Figure 3d). One of the lineages moved west towards the

Ural Mountains and the other eastward. The Ural Mountains have

been suggested as a LGM refugium for several animal species, such

as the field vole (Microtus agrestis) (Jaarola & Searle, 2002). The

MRCA of European bears lived around 25 kya (95% CI: 17–32 kya)

and, as conditions improved, this lineage colonized eastern and

northern Europe, reaching Romania (Figure 3d). The clade that

remained east of the Ural Mountains started to diverge about

25 kya (95% CI: 17–34 kya; Figure 3d). One lineage migrated

towards the Ural Mountains and the other towards Magadan region.

Although the western migration reached the Ural Mountains, this lin-

eage has not been found farther west than the Perm region on the

western slopes of the Urals. As with the 3a1 clade in brown bears,

the common juniper (Juniperus communis) also has a dominant and

widespread clade in Eurasia (Hantemirova et al., 2017), although it

does not reach as far east.

In conclusion, our mitogenome analysis of brown bears in Eurasia

and North America demonstrates that this species shared several

population processes in common with wapiti and modern humans

during the last 50 thousand years, including the Beringian Standstill.

This suggests that wild mammal species and modern humans

responded to the same key limiting factors in the Late Pleistocene.

ACKNOWLEDGEMENTS

The authors thank three anonymous reviewers for their insightful

comments and suggestions, and Dr. Frank Zachos for his generous

help. We also thank Eline Lorenzen for providing sequence data. This

work was supported by the institutional research funding (grant

IUT20-32) from the Estonian Ministry of Education and Research;

BIOGEAST – Biodiversity of East-European and Siberian large mam-

mals on the level of genetic variation in populations, 7th Framework

Programme (contracts 247652 and 2096/7. PR UE/2011/2); Aus-

tralian Research Council; Swedish Environmental Protection Agency;

Norwegian Environment Agency; Research Council of Norway; and

the Finnish Ministry of Agriculture and Forestry.

ORCID

Urmas Saarma http://orcid.org/0000-0002-3222-571X

REFERENCES

Agadjanian, A. K., & Serdyuk, N. V. (2005). The history of mammalian

communities and paleogeography of the Altai Mountains in the Pale-

olithic. Paleontological Journal, 39, 645–821.

Ahn, J., & Brook, E. J. (2014). Siple Dome ice reveals two modes of mil-

lennial CO2 change during the last ice age. Nature Communications, 5,

3723. https://doi.org/10.1038/ncomms4723

Allen, J. R. M., Hickler, T., Singarayer, J. S., Sykesb, M. T., Valdesc, P. J., &

Huntley, B. (2010). Last glacial vegetation of northern Eurasia. Qua-

ternary Science Reviews, 29, 2604–2618. https://doi.org/10.1016/j.

quascirev.2010.05.031

Barnes, I., Matheus, P., Shapiro, B., Jensen, D., & Cooper, A. (2002).

Dynamics of Pleistocene population extinctions in Beringian brown

bears. Science, 295, 2267–2270. https://doi.org/10.1126/science.

1067814

Barr, I. D., & Solomina, O. (2014). Pleistocene and Holocene glacier fluc-

tuations upon the Kamchatka Peninsula. Global and Planetary Change,

113, 110–120. https://doi.org/10.1016/j.gloplacha.2013.08.005

Bidon, T., Janke, A., Fain, S. R., Eiken, H. G., Hagen, S. B., Saarma, U., . . . Hai-

ler, F. (2014). Brown and polar bear Y chromosomes reveal extensive

male-biased gene flow within brother lineages.Molecular Biology and Evo-

lution, 31, 1353–1363. https://doi.org/10.1093/molbev/msu109

Bray, S. C. E. (2010) Mitochondrial DNA analysis of the evolution and

genetic diversity of ancient and extinct bears (PhD Thesis). University

of Adelaide.

Bray, S. C. E., Austin, J. J., Metcalf, J. L., Østbye, K., Østbye, E., Lauritzen,

S.-E, . . . Cooper, A. (2013) Ancient DNA identifies post-glacial

recolonisation, not recent bottlenecks, as the primary driver of con-

temporary mtDNA phylogeography and diversity in Scandinavian

brown bears. Diversity and Distributions, 19, 245–256. https://doi.

org/10.1111/ddi.2013.19.issue-3

Brubaker, L. B., Anderson, P. M., Edwards, M. E., & Lozhkin, A. V. (2005).

Beringia as a glacial refugium for boreal trees and shrubs: New per-

spectives from mapped pollen data. Journal of Biogeography, 32, 833–

848. https://doi.org/10.1111/jbi.2005.32.issue-5

Davison, J., Ho, S. Y. W., Bray, S. C., Korsten, M., Tammeleht, E., Hindrik-

son, M., . . . Saarma, U. (2011) Late-Quaternary biogeographic scenar-

ios for the brown bear (Ursus arctos), a wild mammal model species.

Quaternary Science Reviews, 30, 418–430. https://doi.org/10.1016/j.

quascirev.2010.11.023

Deffontaine, V., Libois, R., Kotl�ık, P., Sommer, R., Nieberding, C., Par-

adis, E., . . . Michaux, J.R. (2005). Beyond the Mediterranean penin-

sulas: Evidence of central European glacial refugia for a temperate

forest mammal species, the bank vole (Clethrionomys glareolus).

Molecular Ecology, 14, 1727–1739. https://doi.org/10.1111/mec.

2005.14.issue-6

402 | ANIJALG ET AL.

Page 10: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

Derenko, M., Malyarchuk, B., Denisova, G., Perkova, M., Litvinov, A.,

Grzybowski, T., & Zakharov, I. (2014). Western Eurasian ancestry in

modern Siberians based on mitogenomic data. BMC Evolutionary Biol-

ogy, 14, 217. https://doi.org/10.1186/s12862-014-0217-9

Drummond, A. J., Suchard, M. A., Xie, D., & Rambaut, A. (2012). Bayesian

phylogenetics with BEAUti and the BEAST 1.7. Molecular Biology and

Evolution, 29, 1969–1973. https://doi.org/10.1093/molbev/mss075

Duchene, S., Duchene, D., Holmes, E. C., & Ho, S. Y. W. (2015). The per-

formance of the date-randomization test in phylogenetic analyses of

time-structured virus data. Molecular Biology and Evolution, 32, 1895–

1906. https://doi.org/10.1093/molbev/msv056

Dulik, M. C., Zhadanov, S. I., Osipova, L. P., Askapuli, A., Gau, L., Gokcu-

men, O., . . . Schurr, T. G. (2012) Mitochondrial DNA and Y chromo-

some variation provides evidence for a recent common ancestry

between native Americans and indigenous Altaians. American Journal

of Human Genetics, 90, 229–246. https://doi.org/10.1016/j.ajhg.

2011.12.014

Goebel, T. (1999). Pleistocene human colonization of Siberia and peopling

of the Americas: An ecological approach. Evolutionary Anthropology, 8,

208–227. https://doi.org/10.1002/(ISSN)1520-6505

Guskov, V. Y., Sheremeteva, I. N., Seredkin, I. V., & Kryukov, A. P. (2013).

Mitochondrial cytochrome b gene variation in brown bear (Ursus arc-

tos Linnaeus, 1758) from southern part of Russian Far East. Russian

Journal of Genetics, 49, 1213–1218. https://doi.org/10.1134/

S1022795413110070

Guthrie, R. D. (2001). Origin and causes of the mammoth steppe: A story

of cloud cover, woolly mammal tooth pits, buckles, and inside-out

Beringia. Quaternary Science Reviews, 20, 549–574. https://doi.org/

10.1016/S0277-3791(00)00099-8

Hantemirova, E. V., Heinze, B., Knyazeva, S. G., Musaev, A. M., Lascoux,

M., & Semerikov, V. L. (2017). A new Eurasian phylogeographical

paradigm? Limited contribution of southern populations to the recol-

onization of high latitude populations in Juniperus communis L.

(Cupressaceae). Journal of Biogeography, 44, 271–282. https://doi.

org/10.1111/jbi.2017.44.issue-2

Hewitt, G. (2000). The genetic legacy of the Quaternary ice ages. Nature,

405, 907–913. https://doi.org/10.1038/35016000

Hilderbrand, G. V., Schwartz, C. C., Robbins, C. T., Jacoby, M. E., Hanley,

T. A., Arthur, S. M., & Servheen, C. (1999). The importance of meat,

particularly salmon, to body size, population productivity, and conser-

vation of North American brown bears. Canadian Journal of Zoology,

77, 132–138. https://doi.org/10.1139/z98-195

Hindrikson, M., Remm, J., Pilot, M., Godinho, R., Stronen, A.V., & Bal-

trunaite, L., . . . Saarma, U. (2017) Wolf population genetics in Europe:

A systematic review, meta-analysis and suggestions for conservation

and management. Biological Reviews, 92, 1601–1629. https://doi.org/

10.1111/brv.12298

Hirata, D., Abramov, A. V., Baryshnikov, G. F., & Masuda, R. (2014). Mito-

chondrial DNA haplogrouping of the brown bear, Ursus arctos (Car-

nivora: Ursidae) in Asia, based on a newly developed APLP analysis.

Biological Journal of the Linnean Society, 111, 627–635. https://doi.

org/10.1111/bij.2014.111.issue-3

Hirata, D., Mano, T., Abramov, A. V., Baryshnikov, G. F., Kosintsev, P. A.,

Vorobiev, A. A., . . . Masuda, R. (2013) Molecular phylogeography of

the brown bear (Ursus arctos) in Northeastern Asia based on analyses

of complete mitochondrial DNA sequences. Molecular Biology and

Evolution, 30, 1644–1652. https://doi.org/10.1093/molbev/mst077

Ho, S. Y. W., Lanfear, R., Bromham, L., Phillips, M. J., Soubrier, J.,

Rodrigo, A. G., & Cooper, A. (2011). Time-dependent rates of molecu-

lar evolution. Molecular Ecology, 20, 3087–3101. https://doi.org/10.

1111/j.1365-294X.2011.05178.x

Ho, S. Y. W., Saarma, U., Barnett, R., Haile, J., & Shapiro, B. (2008). The

effect of inappropriate calibration: Three case studies in molecular

ecology. PLoS ONE, 3, e1615. https://doi.org/10.1371/journal.pone.

0001615

Hoffecker, J. F., Elias, S. A., & O’Rourke, D. H. (2014). Out of Beringia?

Science, 343, 979–980. https://doi.org/10.1126/science.1250768

Hofreiter, M., Serre, D., Rohland, N., Rabeder, G., Nagel, D., Conard, N.,

. . . P€a€abo, S. (2004) Lack of phylogeography in European mammals

before the last glaciation. Proceedings of the National Academy of

Sciences of the United States of America, 101, 12963–12968. https://d

oi.org/10.1073/pnas.0403618101

Hofreiter, M., & Stewart, J. (2009). Ecological change, range fluctuations

and population dynamics during the Pleistocene. Current Biology, 19,

R584–R594. https://doi.org/10.1016/j.cub.2009.06.030

Hoogakker, B. A. A., Smith, R. S., Singarayer, J. S., Marchant, R., Prentice,

I. C., Allen, J. R. M., . . . Tzedakis, C. (2016) Terrestrial biosphere

changes over the last 120 kyr. Climate of the Past, 12, 51–73.

https://doi.org/10.5194/cp-12-51-2016

Huntley, B., Allen, J. R. M., Collingham, Y. C., Hickler, T., Lister, A. M.,

Singarayer, J., . . . Valdes, P.J. (2013) Millennial climatic fluctuations

are key to the structure of Last Glacial Ecosystems. PLoS ONE, 8,

e61963. https://doi.org/10.1371/journal.pone.0061963

Jaarola, M., & Searle, J. B. (2002). Phylogeography of field voles (Microtus

agrestis) in Eurasia inferred from mitochondrial DNA sequences.

Molecular Ecology, 11, 2613–2621. https://doi.org/10.1046/j.1365-

294X.2002.01639.x

Keis, M., Remm, J., Ho, S. Y. W., Davison, J., Tammeleht, E., Tumanov, I.

L., . . . Saarma, U. (2013) Complete mitochondrial genomes and a

novel spatial genetic method reveal cryptic phylogeographical struc-

ture and migration patterns among brown bears in north-western

Eurasia. Journal of Biogeography, 40, 915–927. https://doi.org/10.

1111/jbi.2013.40.issue-5

Kindler, P., Guillevic, M., Baumgartner, M., Schwander, J., Landais, A., &

Leuenberger, M. (2014). Temperature reconstruction from 10 to 120

kyr b2k from the NGRIP ice core. Climate of the Past, 10, 887–902.

https://doi.org/10.5194/cp-10-887-2014

Koblm€uller, S., Vil�a, C., Lorente-Galdos, B., Dabad, M., Ramirez, O., Mar-

ques-Bonet, T., . . . Leonard, J. (2016) Whole mitochondrial genomes

illuminate ancient intercontinental dispersals of grey wolves (Canis

lupus). Journal of Biogeography, 43, 1728–1738. https://doi.org/10.

1111/jbi.12765

Kohli, B. A., Fedorov, V. B., Waltari, E., & Cook, J. A. (2015). Phylogeog-

raphy of a Holarctic rodent (Myodes rutilus): Testing high-latitude bio-

geographical hypotheses and the dynamics of range shifts. Journal of

Biogeography, 42, 377–389. https://doi.org/10.1111/jbi.12433

Korsten, M., Ho, S. Y. W., Davison, J., P€ahn, B., Vulla, E., Roht, M., . . .

Saarma, U. (2009) Sudden expansion of a single brown bear maternal

lineage across northern continental Eurasia after the last ice age: A

general demographic model for mammals? Molecular Ecology, 18,

1963–1979. https://doi.org/10.1111/j.1365-294X.2009.04163.x

Kotl�ık, P., Deffontaine, V., Mascheretti, S., Zima, J., Michaux, J. R., &

Searle, J. B. (2006). A northern glacial refugium for bank voles

(Clethrionomys glareolus). Proceedings of the National Academy of

Sciences of the United States of America, 103, 14860–14864. https://d

oi.org/10.1073/pnas.0603237103

Kurt�en, B. (1968). Pleistocene mammals of Europe. London: Weidenfeld &

Nicolson.

Lambeck, K., Yokoyama, Y., & Purcell, T. (2002). Into and out of the Last

Glacial Maximum: Sea-level change during Oxygen Isotope Stages 3

and 2. Quaternary Science Reviews, 21, 343–360. https://doi.org/10.

1016/S0277-3791(01)00071-3

Lanfear, R., Calcott, B., Ho, S. Y. W., & Guindon, S. (2012). PartitionFin-

der: Combined selection of partitioning schemes and substitution

models for phylogenetic analyses. Molecular Biology and Evolution, 29,

1695–1701. https://doi.org/10.1093/molbev/mss020

Leonard, J. A., Wayne, R. K., & Cooper, A. (2000). Population genetics of

Ice Age brown bears. Proceedings of the National Academy of Sciences

of the United States of America, 97, 1651–1654. https://doi.org/10.

1073/pnas.040453097

ANIJALG ET AL. | 403

Page 11: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

Liu, S., Lorenzen, E. D., Fumagalli, M., Li, B., Harris, K., Xiong, Z., . . .

Wang, J. (2014). Population genomics reveal recent speciation and

rapid evolutionary adaptation in polar bears. Cell, 157, 785–794.

https://doi.org/10.1016/j.cell.2014.03.054

Llamas, B., Fehren-Schmitz, L., Valverde, G., Soubrier, J., Mallick, S., Roh-

land, N., . . . Haak, W. (2016). Ancient mitochondrial DNA provides

high-resolution time scale of the peopling of the Americas. Science

Advances, 2: e1501385. https://doi.org/10.1126/sciadv.1501385

McCarthy, T. M., Waits, L. P., & Mijiddorj, B. (2009). Status of the Gobi

bear in Mongolia as determined by noninvasive genetic methods.

Ursus, 20, 30–38. https://doi.org/10.2192/07GR013R.1

McDevitt, A. D., Zub, K., Kawałko, A., Oliver, M. K., Herman, J. S., &

Wojcik, J. M. (2012). Climate and refugial origin influence the mito-

chondrial lineage distribution of weasels (Mustela nivalis) in a phylo-

geographic suture zone. Biological Journal of the Linnean Society, 106,

57–69. https://doi.org/10.1111/bij.2012.106.issue-1

Meiri, M., Lister, A. M., Collins, M. J., Tuross, N., Goebel, T., Blockley, S.,

. . . Barnes, I. (2014) Faunal record identifies Bering isthmus condi-

tions as constraint to end-Pleistocene migration to the New World.

Proceedings of the Royal Society B: Biological Sciences, 281, 20132167.

Molak, M., Lorenzen, E. D., Shapiro, B., & Ho, S. Y. W. (2013). Phyloge-

netic estimation of timescales using ancient DNA: The effects of

temporal sampling scheme and uncertainty in sample ages. Molecular

Biology and Evolution, 30, 253–262. https://doi.org/10.1093/molbev/

mss232

Murtskhvaladze, M., Gavashelishvili, A., & Tarkhnishvili, D. (2010). Geo-

graphic and genetic boundaries of brown bear (Ursus arctos) popula-

tion in the Caucasus. Molecular Ecology, 19, 1829–1841. https://doi.

org/10.1111/mec.2010.19.issue-9

Pavelkov�a �Ri�c�ankov�a, V., Robovsk�y, J., & Riegert, J. (2014). Ecological

structure of recent and Last Glacial mammalian faunas in Northern

Eurasia: The case of Altai-Sayan refugium. PLoS ONE, 9, e85056.

https://doi.org/10.1371/journal.pone.0085056

Pavelkov�a �Ri�c�ankov�a, V., Robovsk�y, J., Riegert, J., & Zrzav�y, J. (2015).

Regional patterns of postglacial changes in the Palearctic mammalian

diversity indicate retreat to Siberian steppes rather than extinction.

Scientific Reports, 5, 12682.

Pitulko, V. V., Nikolsky, P. A., Girya, E. Y., Basilyan, A. E., Tumskoy, V. E.,

Koulakov, S. A., . . . Anisimov, M. A. (2004) The Yana RHS site:

Humans in the Arctic before the Last Glacial Maximum. Science, 303,

52–56. https://doi.org/10.1126/science.1085219

Rambaut, A., Lam, T. T., Carvalho, L. M., & Pybus, O. G. (2016) Exploring

the temporal structure of heterochronous sequences using TempEst

(formerly Path-O-Gen). Virus Evolution, 2, vew007. https://doi.org/10.

1093/ve/vew007

Ramsden, C., Holmes, E. C., & Charleston, M. A. (2009). Hantavirus evo-

lution in relation to its rodent and insectivore hosts: No evidence for

codivergence. Molecular Biology and Evolution, 26, 143–153.

Ray, N., & Adams, J. M. (2001). A GIS-based vegetation map of the

World at the Last Glacial maximum (25,000–15,000 BP). Internet

Archaeology, 11, http://intarch.ac.uk/journal/issue11/rayadams_toc.

htm)

Saarma, U., Ho, S. Y. W., Pybus, O. G., Kaljuste, M., Tumanov, I. L., Kojola,

I., . . . R~okov, A. M. (2007) Mitogenetic structure of brown bears (Ursus

arctos L.) in northeastern Europe and a new time frame for the forma-

tion of European brown bear lineages. Molecular Ecology, 16, 401–413.

https://doi.org/10.1111/j.1365-294X.2006.03130.x

Saarma, U., & Kojola, I. (2007). Matrilineal genetic structure of the brown

bear population in Finland. Ursus, 18, 30–37. https://doi.org/10.

2192/1537-6176(2007)18[30:MGSOTB]2.0.CO;2

Salomashkina, V. V., Kholodova, M. V., Tutenkov, O. Y., Moskvitina, N. S.,

& Erokhin, N. G. (2014). New data on the phylogeography and

genetic diversity of the brown bear Ursus arctos Linnaeus, 1758 of

Northeastern Eurasia (mtDNA control region polymorphism analysis).

Biology Bulletin, 41, 38–46. https://doi.org/10.1134/S10623590140

10087

Schmitt, T., & Varga, Z. (2012). Extra-Mediterranean refugia: The rule and

not the exception? Frontiers in Zoology, 9, 22. https://doi.org/10.

1186/1742-9994-9-22

Schregel, J., Eiken, H. G., Grøndahl, F. A., Hailer, F., Aspi, J., Kojola, I, . . .

Hagen, S. B. (2015) Y chromosome haplotype distribution of brown

bears (Ursus arctos) in Northern Europe provides insight into popula-

tion history and recovery. Molecular Ecology, 24, 6041–6060.

https://doi.org/10.1111/mec.13448

Servheen, C., Herrero, S., & Peyton, B. (1999). Bears: Status survey and

conservation action plan. Gland: IUCN.

Sommer, R. S., & Benecke, N. (2005). The recolonization of Europe by

brown bears Ursus arctos Linnaeus, 1758 after the Last Glacial Maxi-

mum. Mammal Review, 35, 156–164. https://doi.org/10.1111/mam.

2005.35.issue-2

Stamatakis, A. (2014). RAxML version 8: A tool for phylogenetic analysis

and post-analysis of large phylogenies. Bioinformatics, 30, 1312–

1313. https://doi.org/10.1093/bioinformatics/btu033

Stojak, J., McDevitt, A. D., Herman, J. S., Searle, J. B., & W�ojcik, J. M.

(2015). Post-glacial colonization of eastern Europe from the Car-

pathian refugium: Evidence from mitochondrial DNA of the common

vole Microtus arvalis. Biological Journal of the Linnean Society, 115,

927–939. https://doi.org/10.1111/bij.12535

Svendsen, J. I., Heggen, H. P., Hufthammer, A. K., Mangeruda, J., Pav-

lovc, P., & Roebroeksd, W. (2010). Geo-archaeological investigations

of Palaeolithic sites along the Ural Mountains – On the northern

presence of humans during the last Ice Age. Quaternary Science

Reviews, 29, 3138–3156. https://doi.org/10.1016/j.quascirev.2010.

06.043

Swenson, J. E., Taberlet, P., & Bellemain, E. (2011). Genetics and conser-

vation of European brown bears Ursus arctos. Mammal Review, 41,

87–98. https://doi.org/10.1111/mam.2011.41.issue-2

Taberlet, P., & Bouvet, J. (1994). Mitochondrial DNA polymorphism, phy-

logeography, and conservation genetics of the brown bear Ursus arc-

tos in Europe. Proceedings of the Royal Society B: Biological Sciences,

255, 195–200. https://doi.org/10.1098/rspb.1994.0028

Taberlet, P., Fumagalli, L., Wust-Saucy, A. G., & Cosson, J. F. (1998).

Comparative phylogeography and postglacial colonization routes in

Europe. Molecular Ecology, 7, 453–464. https://doi.org/10.1046/j.

1365-294x.1998.00289.x

Talbot, S. L., & Shields, G. F. (1996). Phylogeography of brown bears

(Ursus arctos) of Alaska and paraphyly within the Ursidae. Molecular

Phylogenetics and Evolution, 5, 477–494. https://doi.org/10.1006/mpe

v.1996.0044

Tamm, E., Kivisild, T., Reidla, M., Metspalu, M., Smith, D. G., Mulligan, C.

J., . . . Malhi, R. S. (2007). Beringian standstill and spread of native

American founders. PLoS ONE, 2, e829. https://doi.org/10.1371/jour

nal.pone.0000829

Tammeleht, E., Remm, J., Korsten, M., Davison, J., Tumanov, I., Saveljev,

A., . . . Saarma, U. (2010) Genetic structure in large, continuous mam-

mal populations: The example of brown bears in northwestern Eura-

sia. Molecular Ecology, 19, 5359–5370. https://doi.org/10.1111/mec.

2010.19.issue-24

To, T.-H., Jung, M., Lycett, S., & Gascuel, O. (2016). Fast dating using

least-squares criteria and algorithms. Systematic Biology, 65, 82–97.

https://doi.org/10.1093/sysbio/syv068

Vulla, E., Hobson, K. A., Korsten, M., Leht, M., Martin, A.-J., Lind, A., &

Saarma, U. (2009). Carnivory is positively correlated with latitude

among omnivorous mammals: Evidence from brown bears, badgers

and pine martens. Annales Zoologici Fennici, 46, 395–415. https://doi.

org/10.5735/086.046.0601

Waits, L. P., Talbot, S. L., Ward, R. H., & Shields, G. F. (1998). Mitochon-

drial DNA phylogeography of the North American brown bear and

404 | ANIJALG ET AL.

Page 12: Large‐scale migrations of brown bears in Eurasia …bearproject.info/wp-content/uploads/2018/11/2018-A245...nization of North America. We sequenced the complete mitochon-drial genomes

implications for conservation. Conservation Biology, 12, 408–417.

https://doi.org/10.1046/j.1523-1739.1998.96351.x

Watson, T. (2017). News Feature: Is theory about peopling of the Ameri-

cas a bridge too far? Proceedings of the National Academy of Sciences

of the United States of America, 114, 5554–5557. https://doi.org/10.

1073/pnas.1705966114

Willerslev, E., Davison, J., Moora, M., Zobel, M., Coissac, E., Edwards, M.

E., . . . Taberlet, P. (2014). Fifty thousand years of Arctic vegetation

and megafaunal diet. Nature, 506, 47–51. https://doi.org/10.

1038/nature12921

Xie, W., Lewis, P. O., Fan, Y., Kuo, L., & Chen, M. H. (2011). Improving

marginal likelihood estimation for Bayesian phylogenetic model selec-

tion. Systematic Biology, 60, 150–160. https://doi.org/10.1093/sysb

io/syq085

BIOSKETCH

Peeter Anijalg is interested in brown bear phylogeography and

population genetics, and the impact of climate change on ancient

and contemporary animal population processes. The remaining

authors have broad interests in molecular ecology, phylogeogra-

phy and biogeography, particularly relating to animals of the Late

Pleistocene and Holocene.

Author contributions: U.S. and P.A conceived the original idea;

all authors collected the data; P.A., S.Y.W.H., K.B., J.D. and U.S.

analysed the data; P.A., S.Y.W.H., J.D. and U.S. wrote the first

draft of the manuscript and all authors contributed to writing

the final version.

SUPPORTING INFORMATION

Additional Supporting Information may be found online in the sup-

porting information tab for this article.

How to cite this article: Anijalg P, Ho SYW, Davison J, et al.

Large-scale migrations of brown bears in Eurasia and to

North America during the Late Pleistocene. J Biogeogr.

2018;45:394–405. https://doi.org/10.1111/jbi.13126

ANIJALG ET AL. | 405