ArticlePDF Available

Genome sequence, population history, and pelage genetics of the endangered African wild dog (Lycaon pictus)

Authors:

Abstract and Figures

Background The African wild dog (Lycaon pictus) is an endangered African canid threatened by severe habitat fragmentation, human-wildlife conflict, and infectious disease. A highly specialized carnivore, it is distinguished by its social structure, dental morphology, absence of dewclaws, and colorful pelage. ResultsWe sequenced the genomes of two individuals from populations representing two distinct ecological histories (Laikipia County, Kenya and KwaZulu-Natal Province, South Africa). We reconstructed population demographic histories for the two individuals and scanned the genomes for evidence of selection. Conclusions We show that the African wild dog has undergone at least two effective population size reductions in the last 1,000,000 years. We found evidence of Lycaon individual-specific regions of low diversity, suggestive of inbreeding or population-specific selection. Further research is needed to clarify whether these population reductions and low diversity regions are characteristic of the species as a whole. We documented positive selection on the Lycaon mitochondrial genome. Finally, we identified several candidate genes (ASIP, MITF, MLPH, PMEL) that may play a role in the characteristic Lycaon pelage.
Content may be subject to copyright.
R E S E A R C H A R T I C L E Open Access
Genome sequence, population history, and
pelage genetics of the endangered African
wild dog (Lycaon pictus)
Michael G. Campana
1,2*
, Lillian D. Parker
1,2,3
, Melissa T. R. Hawkins
1,2,3
, Hillary S. Young
4
, Kristofer M. Helgen
2,3
,
Micaela Szykman Gunther
5
, Rosie Woodroffe
6
, Jesús E. Maldonado
1,2
and Robert C. Fleischer
1
Abstract
Background: The African wild dog (Lycaon pictus) is an endangered African canid threatened by severe habitat
fragmentation, human-wildlife conflict, and infectious disease. A highly specialized carnivore, it is distinguished by
its social structure, dental morphology, absence of dewclaws, and colorful pelage.
Results: We sequenced the genomes of two individuals from populations representing two distinct ecological histories
(Laikipia County, Kenya and KwaZulu-Natal Province, South Africa). We reconstructed population demographic histories
for the two individuals and scanned the genomes for evidence of selection.
Conclusions: We show that the African wild dog has undergone at least two effective population size reductions in
the last 1,000,000 years. We found evidence of Lycaon individual-specific regions of low diversity, suggestive of
inbreeding or population-specific selection. Further research is needed to clarify whether these population reductions
and low diversity regions are characteristic of the species as a whole. We documented positive selection on the Lycaon
mitochondrial genome. Finally, we identified several candidate genes (ASIP,MITF,MLPH,PMEL) that may play a role in
the characteristic Lycaon pelage.
Keywords: Lycaon pictus, Genome, Population history, Selection, Pelage
Background
The African wild dog (Lycaon pictus) is an endangered
canid species (International Union for Conservation of
Nature Red List Classification: C2a (i)) [1]. While the
species formerly ranged over most of sub-Saharan
Africa, wild dogs suffer from a suite of threats including
severe habitat fragmentation, human persecution, and
disease epidemics. They are now restricted to less than
seven percent of their former range [2], with only small,
and frequently declining, remnant populations in frag-
mented pockets of eastern and southern Africa (Fig. 1).
They maintain enormous home ranges (varying between
200 and 2000 km
2
) and naturally live at very low
densities, even compared to other carnivores [3].
Primarily a hunter of antelopes, the African wild dog is a
highly distinct canine. Wild dogs are differentiated from
other canine species by their anatomical adaptations re-
lated to hypercarnivory and cursorial hunting, including
high-crowned, sectorial teeth and the lack of dewclaws
[4]. They have a highly specialized social structure in
which both males and females disperse to form new
packs and only a single dominant pair in each pack re-
produces [5]. Wild dogs are also noted for their colorful
pelage, from which they derive their species name pictus
(painted), and the absence of an undercoat.
Eastern and southern populations of wild dogs are
genetically and morphologically distinct [6], although
there is a large admixture zone covering Botswana,
south-eastern Tanzania, and Zimbabwe [7]. Gene flow
occurs across the speciesentire range [8], which is
unsurprising given the excellent dispersal capabilities
documented in Lycaon pictus [1]. Nevertheless, Marsden
et al. [2] documented both genetic structuring between
extant Lycaon populations (likely the result of habitat
* Correspondence: CampanaM@si.edu
1
Center for Conservation Genomics, Smithsonian Conservation Biology
Institute, 3001 Connecticut Avenue NW, Washington, DC 20008, USA
2
Department of Environmental Science and Policy, George Mason University,
4400 University Drive, Fairfax, VA 22030, USA
Full list of author information is available at the end of the article
© The Author(s). 2016 Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and
reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to
the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver
(http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.
Campana et al. BMC Genomics (2016) 17:1013
DOI 10.1186/s12864-016-3368-9
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
fragmentation) and a recent reduction in effective
population size (N
e
) since the 1980s. Furthermore,
African wild dogs exhibit very little major histocom-
patibility complex variation, which may reflect popu-
lation decline [8].
To better understand African wild dog genomic evolu-
tion, genetic variation and population history, we
shotgun sequenced whole genomes from two individuals
from widely separated populations with very different
modern ecological histories: Laikipia County, Kenya and
Hluhluwe-Imfolozi Park, KwaZulu-Natal Province,
South Africa. Wild dogs disappeared from Laikipia,
Kenya in the 1980s and reappeared in 2000 likely
recolonizing from a small population in neighboring
Samburu district (~50 km distant). The population in
Laikipia alone now numbers more than 150 dogs and 15
packs [9]. We sequenced a female (sampled July 2003)
from this recolonized population. In contrast, wild dogs
were reintroduced to Hluhluwe-Imfolozi Park,
KwaZulu-Natal Province in 1980 and remained as a
single pack for many years [10]. Eventually, Lycaon
breeding effectively ceased until new animals were intro-
duced in 1997 and afterwards (2001, 2003, etc.) [11].
The KwaZulu-Natal wild dogs are now managed as part
of the South African Lycaon metapopulation[12, 13].
We sequenced a male (sampled October 2007), born in
KwaZulu-Natal to parents that were translocated from
Limpopo province in 2003. Therefore, the South African
individuals ancestry represents genes from the northeastern
part of the country.
The genomes from these two populations represent
some of the first published wild canid genomes and are
particularly valuable given the susceptibility of wild dogs
to diseases and habitat fragmentation [9, 14]. We used
our novel genome sequences to reconstruct the last
1,000,000 years of Lycaon genome demography and
population history. We identified over a million poly-
morphic sequence variants for further population-level
study. These variants produced ~35 million predicted
genic effects. We identified over 15,000 candidate genes
that may have undergone adaptation since the Lycaon/
Canis divergence. We found evidence of positive selec-
tion on the Lycaon mitochondrial genome. Finally, we
examined genes involved in canid coat phenotype to
identify candidate genes underlying the characteristic
Lycaon pelage.
Results and discussion
Genome sequencing of the African wild dog
Based on alignment with the domestic dog genome, we
have sequenced ~90% (5.8× mean read depth) of the
Kenyan individuals genome and ~93% (5.7×) of the
South African Lycaon individuals genome. We identified
16,967,383 autosomal sequence variants (including
14,360,480 single nucleotide polymorphisms [SNPs] and
2,606,903 indels) separating our Lycaon genomes from
the domestic dog (Canis familiaris) genome [GenBank:
CanFam3.1] (Additional files 1 and 2) [15]. Of these,
1,092,450 (781,329 SNPs and 311,121 indels) were
polymorphic in the African wild dog. The remaining
15,874,933 autosomal sequence variants (13,579,151
SNPs and 2,295,782 indels) were monomorphic in the
two Lycaon individuals. We identified 717,870 X-
chromosomal variants (619,606 SNPs and 98,264 indels),
of which 32,801 (23,001 SNPs and 9,800 indels) were
polymorphic and 685,069 (596,605 SNPs and 88,464
indels) were monomorphic in the African wild dog.
Additionally, we sequenced the Kenyan and South
African wild dog mitochondrial genomes to depths of
943× and 1021×, respectively. We annotated the Lycaon
Fig. 1 South African wild dog pack (top)andmapofextantandformer
wild dog range (bottom). The sampling locations of the two individuals
are noted on the map. Ranges are modified from Woodroffe
and Sillero-Zubiri [1] and Marsden et al. [2]. Extant range data
used with permission from the International Union for the Conservation
of Nature [Woodroffe R, Sillero-Zubiri C 2012. Lycaon pictus.In:IUCN
2016. IUCN Red List of Threatened Species. Version 20162. http://
www.iucnredlist.org. Downloaded 12 July 2016]. Photograph by
Micaela Szykman Gunther
Campana et al. BMC Genomics (2016) 17:1013 Page 2 of 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
autosomal and X-chromosomal sequences using the do-
mestic dog genome annotations [GenBank:CanFam3.1.81]
and the mitochondrial genome [GenBank:NC_002008.4]
and MSY sequences [GenBank:KP081776.1] using their
reference sequence annotations [1517].
Lycaon demographic history
We analyzed the two wild dogsautosomal population
histories using PSMC (Fig. 2) [18]. Both genomes exhib-
ited a strong reduction in N
e
starting 700,000 years ago
from maximum N
e
s of ~28,000 (Kenyan) and ~35,000
(South African) individuals and leveling off 200,000 years
ago at N
e
s of ~7,000 individuals each. Analysis of the
Kenyan individuals X - chromosomal history using
PSMC showed a similar pattern (Fig. 2). N
e
fell from a
maximum of ~40,000 individuals 600,000 years ago to
~7000 individuals 200,000 years ago. This N
e
reduction
may represent a past population bottleneck e.g. [19] or
lineage splitting e.g. [20]. Further genomic analysis of in-
dividuals from across the Lycaon range, especially those
from larger founder populations, would help clarify the
cause of this pattern.
PSMC analysis of the Kenyan X chromosome revealed
a secondary reduction in N
e
started 70,000 years ago to
a final N
e
of ~2000 individuals 10,000 years ago. We are
unable to infer more recent population history accur-
ately using this method due to the limited numbers of
available mutations in the short time frame [18]. Further
research using historical museum Lycaon specimens
would fill in this temporal gap e.g. [21]. While our
reconstructions were robust to coverage variation (see
Demographic history reconstructionbelow), higher
coverage genomes would also increase the resolution of
population history reconstructions [20].
Candidate selected processes and genomic regions
Based on the domestic dog genome annotations, SnpEff
4.1 L predicted that the identified Lycaon autosomal and
X-chromosomal sequence variants would cause 34,001,288
and 36,362,161 genic effects in the Kenyan and South
African individuals, respectively (Additional files 1 and 2)
[22]. The majority of sequence variant effects fell within
introns (Kenyan: 62.604%, South African: 63.383%) and
intergenic regions (Kenyan: 23.081%, South African:
23.162%). To discover candidate genes that have diverged
since Lycaon s divergence with Canis, we identified Lycao n
genes that contained missensemutationsandstopcodon
gains using SnpSift 4.1 L [23]. We identified 15,611 (15,565
genes with missense mutations and 799 with stop codon
gains) Kenyan and 9793 (9440 genes containing missense
mutations and 741 with stop codon gains) South African
wild dog candidate divergent genes. 9506 divergent genes
(9159 with missense mutations and 653 with stop codon
gains) were found in both wild dogs. These divergent genes
were determined by 47,059 Kenyan sequence variants
(sequenced at 8.6× mean coverage) of and 27,893 South
African variants (6.0× mean coverage). 25,149 variants were
shared between the two Lycao n individuals. These variants
were very homozygous (Kenyan: 95%, South African: 95%),
whichsuggeststhattheyaretheresultofdivergence
between the Lycao n and Canis clades, rather than more
recent variants arising within Lycaon.
We annotated the candidate divergent genesfunctions
using DAVID 6.7 with domestic dog (option Canis
lupus) as the genomic background [24]. We found 76
and 29 enriched processes in the Kenyan and South
African individuals, respectively (Additional files 3 and
4). We filtered these terms with a Benjamini-Hochberg
false discovery rate of 0.05 [25]. After filtration, seven
terms (Complement and coagulation cascades,ECM-
0
1
2
3
4
5
6
104105106
Effective population size (x104)
Years before present
South Africa (Autosomes)
Kenya (Autosomes)
Kenya (X Chromosome)
Fig. 2 Reconstruction of the Lycaon individualsautosomal and X-chromosomal demographic history using the pairwise sequentially Markovian
coalescent. Initial results are plotted using dark-colored curves, with the bootstrap replicates plotted in lighter hues of the corresponding colors
Campana et al. BMC Genomics (2016) 17:1013 Page 3 of 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
receptor interaction,Neuroactive ligand-receptor inter-
action,Hematopoietic cell lineage,Lysosome,ABC
transporters,Aminoacyl-tRNA biosynthesis) were sig-
nificantly enriched in the Kenyan individual. Olfactory
transductionwas significant in the South African indi-
vidual. The differences in significant terms may reflect
population-specific selection pressures on wild dogs.
Further population-level investigation is needed to deter-
mine the roles these pathways play in Lycaon evolution.
In order to identify regions of low and high diversity, we
calculated the numbers of segregating SNP sites across the
Lycaon autosomes in 100,000 bp non-overlapping windows
using VCFtools 0.1.15 (Fig. 3) [26, 27]. By averaging over
large genomic windows, we limited the effects of sequen-
cing errors and allelic drop-out due to low sequencing
coverage. We identified 768,416 segregating Lycaon SNPs
(Kenyan: 398,891 SNPs, South African: 434,911 SNPs). We
observed individual-specific regions of low diversity (<10
segregating SNPs/100,000 bp). The Kenyan individual had
runs of low diversity (contiguous regions of low diversity at
least 5 million bp long) on chromosomes 4, 6, 7, 12, 15, 21,
27, and 30, while the South African individual had runs of
low diversity on chromosomes 1, 5, 8, 12, 14, 19, 27, 29, 30,
34, 36 and 38. These low-diversity regions may be the result
of inbreeding and/or population-specific natural selection.
Due to the long lengths of these low-diversity runs, encom-
passing numerous genes, we are not currently able to link
low diversity levels to selection on individual genes. These
results are not surprising since both populations are
recently re-established, either by natural recolonization
(Laikipia, Kenya) or deliberate reintroduction (KwaZulu-
Natal, South Africa). Previous genetic investigations using
microsatellites and mitochondrial DNA found some
evidence of rare inbreeding in wild dogs from the Greater
Limpopo Transfrontier Conservation Area and KwaZulu-
Natal [10, 28]. However, free-ranging wild dogs strongly
avoid incestuous matings [10]. For instance, at KwaZulu-
Natal, Becker et al. [10] observed only one of six breeding
pairs being more closely related than third-order kin. While
our chromosomal diversity data do not permit us to discern
Fig. 3 Segregating autosomal SNP sites across the Lycaon genomes. Chromosomes are distinguished by color and separated by black lines.Thenumber
of segregating SNPs per 100,000 bp window is plotted on the y-axis in logarithmic scale. We identified population-specific regions of low diversity inboth
the Kenyan (chromosomes 4, 6, 7, 12, 15, 21, 27, and 30) and South African (chromosomes 1, 5, 8, 12, 14, 19, 27, 29, 30, 34, 36 and 38) individuals. There are
also highly variable regions on chromosomes 3 and 16 in both individuals, chromosome 26 in the Kenyan individual, and chromosome 19 in the South
African individual
Campana et al. BMC Genomics (2016) 17:1013 Page 4 of 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
between inbreeding and/or population-specific natural
selection, the possibility of inbreeding is, therefore, con-
cerning from a conservation standpoint. Additional
population-level data are required to determine the causes
and effects of these low-diversity regions.
We also identified regions of high diversity (>200
segregating SNPs per 100,000 bp) on chromosomes 3
and 16 in both individuals and chromosome 19 in the
South African individual (Fig. 3). The high-diversity re-
gions comprised 0.44% of the total segregating SNPs in
both individuals and included 1138 Kenyan and 970
South African segregating SNPs on chromosome 3,646
Kenyan and 704 South African segregating SNPs on
chromosome 16, and 249 South African segregating
SNPs on chromosome 19. Using the UCSC Genome
Browser [29], we scanned the genome within 100,000 bp
upstream and downstream of these high-diversity
regions for known genes mapped to CanFam3.1. These
included FER,FKBP3,SNRPF,andPJA2 on chromosome
3, HUS1 and PKDIL1 on chromosome 16, and RNF150
and TBC1D9 on chromosome 19. These genic regions
may have undergone positive or relaxed selection
since the Lyc aon/Canis divergence or may represent
chromosomal duplications. Future functional analyses
will determine which roles (if any) these genes play in
Lycaon evolution.
Positive selection on the mitochondrial genome
Each of the 13 protein-encoding mitochondrial genes
from the two novel Lycaon mitochondrial genomes were
aligned against the corresponding genic sequences from
the domestic dog mitochondrial reference sequence and
the one publically available near-complete Lycaon pictus
mitochondrial genome sequence [GenBank:KT448283.1]
[16, 30]. We calculated ratios of non-synonymous substi-
tutions per non-synonymous site to synonymous substi-
tutions per synonymous site (dN/dS) and tested for
positive selection using the codon-based Z-test in
MEGA6 [31]. We found evidence for positive selection
on 12 of 13 genes (COX1,COX2,COX3,ATP6,ATP8,
ND1,ND2,ND3,ND4,ND4L,ND5,CYTB; Additional
file 5). To determine whether selection occurred primar-
ily on the Lycaon or Canis branches, we reran these
analyses excluding the domestic dog sequence. We
found evidence for positive selection on all 13 Lycaon
mitochondrial genes (Additional file 5).
To confirm branch specific signatures of selection, we
aligned the Lycaon CYTB sequences against published
canid taxa [GenBank: AY656746, EU442884, GU063864,
JF342908, KF646248, KT4482734, KU696390,
KU696404, KU696408, NC_008093, NC_013445] using
MAFFT 7.222 [32] as implemented in Geneious 10.0.2
(Biomatters, Ltd., Auckland, New Zealand). We then
generated a phylogenetic tree with FastTree 1.0 [33].
Using the codonmlalgorithm in PAML 4.8 [34], we per-
formed pairwise comparisons on CYTB codon data and
compared the likelihood of the alternate (non-fixed
omega values) or null hypotheses (fixed omega values).
A likelihood ratio test was calculated from ln values
obtained from these comparisons to determine where
evidence of selection was occurring in canids. We found
significant evidence (p0.001) of positive selection
between African wild dogs and coyotes (Canis latrans)
(statistic: 128.88) and between Lycaon and Cuon (statis-
tic: 81.48). We also found branch-specific selection be-
tween the Kenyan and South African Lycaon individuals
(statistic: 92.54).
The 13 candidate selected Lycaon mitochondrial
genes are involved in the electron transport chain and
adenosine triphosphate synthesis. This suggests
natural selection on Lyc aon metabolic processes e.g.
[35, 36], which is likely given their unique antelope
hunting strategies and diet. Moreover, these results
are consistent with African wild dogsvery high meta-
bolic rate and hunting energy expenditure in com-
parison to domestic dogs [37].
Pelage genes
We extracted CDS corresponding to 11 genes involved
in canid coat color (agouti signaling peptide [ASIP],
β-defensin 103 [DEFB103A], melanocortin 1 receptor
[MC1R], melanophilin [MLPH], microphthalmia-associated
transcription factor [MITF], premelanosome protein
[PMEL], tyrosinase-related protein 1 [TYRP1]) and type
(fibroblast growth factor 5 [FGF5], keratin 71 [KRT71],
R-spondin 2 [RSPO2]) [3846] using Geneious 9.0.4. In
cases where there were multiple isoforms or CDS annota-
tions (DEFB103A,FGF5,MITF,PMEL), we chose the
longest variant alignable to the domestic dog CDS reference
sequence to maximize detection power. To detect positively
selected genes, we identified non-synonymous and syn-
onymous SNPs compared to the domestic dog sequence
and calculated the non-synonymous/synonymous (N/S)
ratio (Additional file 6).
ASIP and PMEL had elevated N/Sratios suggestive of
positive selection (5.00 and 9.00 respectively). Lycaon
PMEL also had a stop codon gain at amino acid 341,
suggesting selection at this locus. Additionally, we found
a threonine insertion at amino acid 371 in the Lycaon
MLPH gene and a six amino acid deletion corresponding
to domestic dog amino acids 186191 in the Lycaon
MITF gene. To further characterize these four candidate
genes, we compared the Lycaon coding sequences
against all publically available canid sequences using
BLAST+ 2.5.0 [47]. None of the Lycaon ASIP,PMEL,
MITF, and MLPH CDS haplotypes have been identified
in other canids previously. Lycaon ASIP shares 99%
nucleotide identity, but only 96% amino acid identity,
Campana et al. BMC Genomics (2016) 17:1013 Page 5 of 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
with domestic dogs. Wild dog PMEL shares both 99%
nucleotide identity and amino acid identity with do-
mestic dogs. Excluding the six amino acid deletion,
Lycaon MITF haplotypes had >99% nucleotide and
amino acid identity with domestic dogs. Lycaon
MLPH haplotypes had 98% nucleotide identity and
9798% amino acid identity (excluding the amino
acid insertion) to domestic dogs.
These four genes are strong candidates to explain the
characteristic Lycaon pelage. Mutations in ASIP alter
relative production of eumelanin and pheomelanin,
resulting in lighter and darker hair colors, in numerous
species including domestic dogs [41]. Variants in PMEL
cause merle patterning in domestic dogs [40]. MITF is
associated with white-spotting phenotypes in domestic
dogs and causes coat color variants in laboratory mice
(Mus musculus). MITF variants are also associated with
deafness, small eye size, and poor bone resorption in
mice [46]. Nevertheless, further laboratory assays (such
as transgenic experiments) are needed to confirm that
these identified variants are functional and to determine
their phenotypic effects. Furthermore, our data do not
permit us to distinguish between positive selection on
the Lycaon and Canis branches (e.g. coat variation
associated with dog domestication and breed development).
Conclusions
We provide two genome sequences of Lycaon pictus,
representing two individuals from highly divergent
ecological regions (Laikipia County, Kenya and
KwaZulu-Natal Province, South Africa). We identified
over a million polymorphic Lycaon SNPs, useful for fur-
ther population-level analyses. Analyses of these ge-
nomes showed that extant Lycaon populations have
endured at least two population contractions within the
last 1,000,000 years. We identified chromosomal regions
of high and low diversity and over 15,000 candidate
divergent genes. Furthermore, Lycaon mitochondrial
genomes have undergone positive selection, suggestive
of selection for metabolic processes. Finally, we identi-
fied four candidate genes (ASIP,MITF,MLPH,PMEL)
that may be involved in Lycaon pelage patterns.
Methods
Samples
We sequenced two Lycaon pictus individuals sampled
during previous studies: a female from Laikipia County,
Kenya [2, 9] and a male from Hluhluwe-Imfolozi Park,
KwaZulu-Natal Province, South Africa [10, 48, 49]. The
Kenyan individual was sampled under an Institutional
Animal Care and Use Committee (IACUC) protocol ap-
proved by the University of California, Davis (10813) [9],
while the South African individual was sampled under
IACUC protocols approved by the Smithsonian National
Zoological Park (08-21) and Humboldt State University
(06/07.W.209.A) [10].
Laboratory methods
DNA was extracted from blood samples using Qiagen
blood and tissue kits (Qiagen, Valencia, CA, USA) and
sheared to ~350 bp using a Q800R sonicator (Qsonica,
LLC, Newtown, CT, USA). Double-indexed Illumina
libraries were built from the sheared DNA using the
KAPA Library Preparation Kit Illumina (Kapa Biosystems,
Wilmington, MA, USA) with purification steps per-
formed using carboxyl paramagnetic beads [50].
Library quality was ensured via fluorometric analysis
using Qubit® dsDNA HS assays (Life Technologies,
Carlsbad, CA, USA), quantitative PCR using the
KAPA Library Quantification Kit Illumina/Universal
(Kapa Biosystems, Wilmington, MA, USA), and ana-
lysis on a 2100 Bioanalyzer (Agilent Technologies,
Santa Clara, CA, USA) high-sensitivity DNA chip.
Libraries were pooled equimolarly and 2 × 250 bp
paired-end sequenced on a HiSeq 2500 lane (Illumina,
Inc., San Diego, CA, USA).
Sequence quality control
Read pairs were demultiplexed using the BaseSpace® pipe-
line (Illumina, Inc., San Diego, CA, USA). 75,651,396 and
63,766,266 read pairs were generated for the Kenyan and
South African Lycaon samples respectively. Raw reads
were trimmed and adapter artifacts were removed using
Trimmomatic 0.33 (options ILLUMINACLIP: NexteraPE-
PE.fa:2:30:10 LEADING:3 TRAILING:3 SLIDINGWIN-
DOW:4:20 MINLEN:36) [51]. Library sequence quality
was confirmed using FastQC 0.11.2 [52].
Mitochondrial genome assembly
The quality-controlled reads were aligned against the
circularized domestic dog reference mitochondrial
genome [GenBank:NC_002008.4] [16] using Geneious
8.1.6 (medium-low sensitivity, 5 alignment iterations,
minimum mapping quality 30). The aligned reads were
merged using FLASH 1.2.11 (option M 250) [53]. PCR
duplicates were removed from the merged reads using
CD-HIT-DUP 0.5 [54]. The deduplicated reads were
then realigned against the dog reference mitochondrial
genome in Geneious 8.1.6 (medium sensitivity align-
ment, 10 alignment iterations, minimum mapping
quality 30) to generate the final sequences.
Autosomal assembly
The non-mitochondrial reads were merged using
FLASH 1.2.11 (option M 250) [51]. The merged,
unmerged, and unpaired reads were then concatenated
and treated as single-end sequences for downstream
processing. The concatenated reads were aligned against
Campana et al. BMC Genomics (2016) 17:1013 Page 6 of 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
the autosomes of the domestic dog genome build
CanFam3.1 [15] using the memalgorithm in BWA
0.7.12 [55, 56]. Aligned reads below mapping quality 30
and PCR duplicates were removed using the view
(option q 30) and rmdup(option s) commands in
SAMtools 1.3 [57]. Sequence variants (minimum quality
20) were identified using the SAMtools 1.3 mpileup
command (option C50) and BCFtools 1.3 call
command (option m) pipeline [57, 58]. Genome com-
pleteness was evaluated using BUSCO 1.1b1 and the
Vertebratesgene set (Additional files 7, 8 and 9) [59].
Genome assemblies were very complete: compared to
the 1663 completeBUSCO autosomal orthologs found
in the domestic dog, 1614 (97%) of the Kenyan and 1665
(100.1%) of the South African individuals orthologs were
complete. Similarly, 747 fragmentedBUSCO autosomal
orthologs were found in the domestic dog, compared to
688 (92.1%) for the Kenyan individual and 707 (94.6%)
for the South African individual.
Allosomal assembly
The Kenyan Lycaon individual was a female, while the
South African individual was a male. To reconstruct the
allosomal sequences, we aligned the unmapped,
concatenated nuclear reads against either the CanFam3.1
X chromosome assembly (Kenyan individual) [15] or
both the CanFam3.1 X chromosome and domestic dog
MSY chromosome assemblies (South African individual)
[17] using the memalgorithm in BWA 0.7.12 [55, 56].
Aligned reads below mapping quality 30 and PCR dupli-
cates were removed using the view(option q 30) and
rmdup(option s) commands in SAMtools 1.3 [57].
Sequence variants (minimum quality 20) were identified
using the SAMtools 1.3 mpileupcommand (option
C50) and BCFtools 1.3 callcommand (option m)
pipeline [57, 58]. The mapped MSY reads were then
realigned against the reference sequence [Gen-
Bank:KP081776.1] using Geneious 8.1.7 (medium sensi-
tivity alignment, five alignment iterations). Y coding
region sequences were extracted based on the domestic
dog MSY annotations, and consensus sequences were
generated using Geneious 8.1.7 (options Highest Quality
and Total). We excluded non-coding regions from
analysis due to the Y chromosomes large number of re-
petitive elements, which complicates accurate alignment
[17]. We identified 87 Y SNPs between the African wild
dog and the domestic dog, of which 32 were silent muta-
tions, 53 produced amino acid substitutions, and two
caused gene truncations (Additional file 10).
Demographic history reconstruction
Autosomal population history parameters were recon-
structed using PSMC r62 (options N25 t15r5 p
64*1, minimum quality 20) and tested with 100
bootstrap replicates [18]. We calculated the autosomal
mutation rate for each Lycaon individual using the total
number of identified autosomal sequence variants
(13,985,381 and 15,132,667 for the Kenyan and South
African individuals respectively), an estimated autosomal
genome size of 2.3 Gbp, an estimated generation time of
5 years/generation [7] and an estimated divergence time
from the Canis/Cuon clade of 2.74 million years ago
(95% highest posterior density: 2.153.38 million years
ago) [30]. We estimated the Kenyan and South African
Lycaon mutation rates as 5.5 × 10
9
mutations/site/gen-
eration (range: 5.07.1 × 10
9
mutations/site/generation)
and 6.0 × 10
9
mutations/site/generation (range: 4.9
7.7 × 10
9
mutations/site/generation) respectively. Final
PSMC demographic reconstructions were scaled based
on an estimated generation time of 5 years/generation
[7] and a mutation rate of 5.8 × 10
9
mutations/site/gen-
eration. While variation of the reconstruction scaling
within the extremes of the estimated mutation rates
(4.97.7 × 10
9
mutations/site/generation) varied
estimates of N
e
and timing of population history events,
overall demographic history patterns remained similar.
The Kenyan X-chromosomal history was recon-
structed separately from the autosomal history. We cal-
culated the X-chromosomal mutation rate for the
Kenyan individual using the total number of observed
Kenyan X-chromosomal variants (634,216), the same di-
vergence and generation times as for the autosomal ana-
lyses and a chromosome size of 124 Mbp. We estimated
the Kenyan X-chromosomal mutation rate as 4.7 × 10
9
mutations/site/generation (range: 3.85.9 × 10
9
muta-
tions/site/generation). We did not estimate the
X-chromosomal mutation rate for the South African
male due to his hemizygosity. X-chromosomal PSMC
demographic reconstruction and scaling parameters
were the same as for the autosomal analyses except
that the results were scaled with a mutation rate of
4.7 × 10
9
mutations/site/generation. Variation of the
mutation rate scaling again did not affect inference of
demographic history.
To test the effects of coverage on our demographic re-
constructions, we repeated the PSMC analyses under
medium depth stringency settings (minimum sequencing
depth 2, maximum sequencing depth 12) recommended
by PSMCs authors and high depth stringency settings
(minimum sequencing depth 10) recommended in [20].
We recovered nearly identical demographic reconstruc-
tions under the medium depth stringency settings
(Additional file 11). Under the high depth stringency
settings, we were unable to resolve the Kenyan
X-chromosomal reconstruction due to missing data.
Nevertheless, our autosomal reconstructions under the
high depth stringency recovered very similar demo-
graphic histories for the last 1,000,000 years, except that
Campana et al. BMC Genomics (2016) 17:1013 Page 7 of 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
N
e
estimates were larger (particularly in the South
African individual) (Additional file 12). Increased N
e
estimates are expected since heterozygotes are more
likely to be observed in the higher coverage regions [20].
Therefore, we conclude that our reconstructed demo-
graphic patterns were robust to coverage variation.
Interestingly, under the high depth stringency
settings, we also reconstructed two additional contrac-
tion and expansion cycles between 10,000,000 and
600,000 years ago in the South African individuals
population history (Additional file 12). These results
require further verification with additional genome
sequences.
Additional files
Additional file 1: Genic effects of SnpEff-annotated autosomal and
X-chromosomal variants for the Kenyan Lycaon individual. (HTML 1596 kb)
Additional file 2: Genic effects of SnpEff-annotated autosomal and
X-chromosomal variants for the South African Lycaon individual.
(HTML 1535 kb)
Additional file 3: DAVID-annotated enriched functional processes for
the Kenyan Lycaon individual. For each functional term, the total gene
count, the associated genes, and the Benjamini-Hochberg false discovery
rate value are given. (XLSX 23 kb)
Additional file 4: DAVID-annotated enriched functional processes for
the South African Lycaon individual. For each functional term, the total
gene count, the associated genes, and the Benjamini-Hochberg false
discovery rate value are given. (XLSX 16 kb)
Additional file 5: Positive selection test results on the Lycaon
mitochondrial genome. Zstatistics are included in parentheses after the
probability for each of the three competing models (Neutrality, Positive
Selection, Purifying Selection). Results for the tests excluding the domestic
dog sequence are denoted by Lycaonin the column heading. (XLSX 48 kb)
Additional file 6: Positive selection scan results on candidate Lycaon
pelage genes. Genes were aligned to the reference CDS provided in the
table. Gene coordinates are given with reference to the domestic dog
genome [GenBank:CanFam3.1]. The total numbers of fixed and polymorphic
synonymous and non-synonymous SNPs and the N/Sratios are provided.
Also listed are non-SNP variants identified during the scan. (XLSX 41 kb)
Additional file 7: Autosomal BUSCO orthologs identified in the
CanFam3.1 assembly. Scaffolds are identified by GenBank accession.
(XLSX 182 kb)
Additional file 8: Autosomal BUSCO orthologs identified in the Kenyan
Lycaon pictus genome. Scaffolds are identified by GenBank accession.
(XLSX 166 kb)
Additional file 9: Autosomal BUSCO orthologs identified in the South
African Lycaon pictus genome. Scaffolds are identified by GenBank
accession. (XLSX 170 kb)
Additional file 10: Genic effects of Lycaon MSY coding sequence
variants. (XLSX 55 kb)
Additional file 11: Reconstruction of the Lycaon individuals
autosomal and X-chromosomal demographic history using the
pairwise sequentially Markovian coalescent with medium coverage
stringency. Initial results are plotted using dark-colored curves, with
the bootstrap replicates plotted in lighter hues of the corresponding
colors. (EPS 252 kb)
Additional file 12: Reconstruction of the Lycaon individualsautosomal
and X-chromosomal demographic history using the pairwise sequentially
Markovian coalescent with high coverage stringency. Initial results are
plotted using dark-colored curves, with the bootstrap replicates plotted
in lighter hues of the corresponding colors. (EPS 189 kb)
Abbreviations
bp: Base pairs; CDS: Coding sequence; dN/dS: Non-synonymous substitutions
per non-synonymous site to synonymous substitutions per synonymous site;
IACUC: Institutional Animal Care and Use Committee; MSY: Male-specific Y
chromosome; N/S: Non-synonymous substitutions to synonymous substitutions;
N
e
: Effective population size; SNP: Single nucleotide polymorphism
Acknowledgements
We thank Penny Becker (United States Fish and Wildlife Service), Michael
Somers (University of Pretoria) and David Wildt (Smithsonian Conservation
Biology Institute) for facilitating collection of the South African sample. We
are grateful for the assistance of Adam Ferguson (Mpala Research Centre) in
the production of the Lycaon range map. We thank the members of the
Center for Conservation Genomics, Smithsonian Conservation Biology
Institute for helpful advice on this manuscript.
Funding
The Morris Animal Foundation (D14ZO-308) and the National Geographic
Society (884610) supported this research.
Availability of data and materials
The data sets supporting the results of this article are available in the
Sequence Read Archive and GenBank repositories, (BioProject PRJNA304992;
http://www.ncbi.nlm.nih.gov/bioproject/PRJNA304992; GenBank
LPRA00000000 and LPRB00000000; http://www.ncbi.nlm.nih.gov/nuccore/
LPRA00000000 and http://www.ncbi.nlm.nih.gov/nuccore/LPRB00000000).
Authorscontributions
MGC performed the molecular genetic assays, carried out bioinformatics
analyses, and wrote the manuscript. KHM and RFC conceived the study.
MTRH and LDP performed bioinformatics analyses and contributed to the
manuscript. RW, MSG, and JEM provided the Lycaon samples. KHM, RFC, HSY,
RW and JEM participated in the design of the study. All authors reviewed
and approved the manuscript.
Competing interests
The authors declare that they have no competing interests.
Consent for publication
Not applicable.
Ethics approval and consent to participate
The Lycaon pictus individuals were sampled under IACUC protocols approved
by the University of California, Davis (10813), the Smithsonian National
Zoological Park (0821), and Humboldt State University (06/07.W.209.A).
Ezemvelo KZN Wildlife, the Kenya Wildlife Service, the National Museums of
Kenya, the Kenya National Council for Science and Technology, and Mpala
Research Centre gave permission and relevant permits to conduct this research
and provided logistical support. All experiments were conducted in compliance
with the Convention on the Trade in Endangered Species of Wild Fauna and
Flora and the International Union for the Conservation of Nature Policy
Statement on Research Involving Species at Risk of Extinction.
Author details
1
Center for Conservation Genomics, Smithsonian Conservation Biology
Institute, 3001 Connecticut Avenue NW, Washington, DC 20008, USA.
2
Department of Environmental Science and Policy, George Mason University,
4400 University Drive, Fairfax, VA 22030, USA.
3
Division of Mammals, National
Museum of Natural History, MRC 108, Smithsonian Institution, Washington,
DC 20013, USA.
4
Department of Ecology, Evolution and Marine Biology,
University of California Santa Barbara, Santa Barbara, CA 93106, USA.
5
Department of Wildlife, Humboldt State University, 1 Harpst St, Arcata, CA
95521, USA.
6
Institute of Zoology, Zoological Society of London, Regents
Park, London NW1 4RY, UK.
Received: 22 July 2016 Accepted: 2 December 2016
References
1. Woodroffe R, Sillero-Zubiri C. Lycaon pictus. The IUCN Red List of
Threatened Species. 2012;2012:e.T12436A16711116.
Campana et al. BMC Genomics (2016) 17:1013 Page 8 of 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2. Marsden CD, Woodroffe R, Mills MGL, McNutt W, Creel S, Groom R, et
al. Spatial and temporal patterns of neutral and adaptive genetic
variation in the endangered African wild dog (Lycaon pictus). Mol Ecol.
2012;21:137993.
3. Woodroffe R, Ginsberg JR. Introduction. In: Woodroffe R, Ginsberg J, Macdonald
D, editors. The African wild dog. Gland and Cambridge: IUCN; 1998. p. 16.
4. Hartstone-Rose A, Werdelin L, De Ruiter DJ, Berger LR, Churchill SE. The
Plio-Pleistocene ancestor of wild dogs, Lycaon sekowei n. sp. J Paleont.
2010;84:299308.
5. Girman DJ, Mills MGL, Geffen E, Wayne RK. A molecular genetic analysis of
social structure, dispersal, and interpack relationships of the African wild
dog (Lycaon pictus). Behav Ecol Sociobiol. 1997;40:18798.
6. Girman DJ, Kat PW, Mills MGL, Ginsberg JR, Borner M, Wilson V, et al.
Molecular genetic and morphological analyses of the African wild dog
(Lycaon pictus). J Hered. 1993;84:4509.
7. Girman DJ, Vilà C, Geffen E, Creel S, Mills MGL, McNutt JW, et al. Patterns of
population subdivision, gene flow and genetic variability in the African wild
dog (Lycaon pictus). Mol Ecol. 2001;10:170323.
8. Marsden CD, Mable BK, Woodroffe R, Rasmussen GSA, Cleaveland S,
McNutt JW, et al. Highly endangered African wild do gs (Lycaon pictus)
lack variation at the major histocompatibility complex. J Hered.
2009;100 Suppl 1:S5465.
9. Woodroffe R. Demography of a recovering African wild dog (Lycaon pictus)
population. J Mammal. 2011;92:30515.
10. Becker PA, Miller PS, Szykman Gunther M, Somers MJ, Wildt DE, Maldonado
JE. Inbreeding avoidance influences the viability of reintroduced
populations of African wild dogs (Lycaon pictus). PLoS ONE. 2012;7, e37181.
11. Gusset M, Slotow R, Somers MJ. Divided we fail: the importance of social
integration for the re-introduction of endangered African wild dogs (Lycaon
pictus). J Zool. 2006;270:50211.
12. Davies-Mostert HT, Mills MGL, Macdonald DW. A critical assessment of
South Africas managed metapopulation recovery strategy for African wild
dogs. In: Hayward MW, Somers MJ, editors. Reintroduction of top-order
predators. London: Wiley-Blackwell; 2009.
13. Lindsey PA, Alexander R, Du Toit JT, Mills MGL. The cost efficiency of wild
dog conservation in South Africa. Conserv Biol. 2005;19:120514.
14. Woodroffe R, Prager KC, Munson L, Conrad PA, Dubovi EJ, Mazet JA.
Contact with domestic dogs increases pathogen exposure in endangered
African wild dogs (Lycaon pictus). PLoS ONE. 2012;7, e30099.
15. Hoeppner MP, Lundquist A, Pirun M, Meadows JRS, Zamani N,
Johnson J, et al. An improved canine genome and a comprehensive
catalogue of coding genes and non-coding transcripts. PLoS ONE.
2014;9, e91172.
16. Kim KS, Lee SE, Joeng HW, Ha JH. The complete nucleotide sequence of the
domestic dog (Canis familaris) mitochondrial genome. Mol Phylogenet Evol.
1998;10:21020.
17. Li G, Davis BW, Raudsepp T, Pearks Wilkerson AJ, Mason VC, Ferguson-Smith M,
et al. Comparative analysis of mammalian Y chromosomes illuminates ancestral
structure and lineage-specific evolution. Genome Res. 2013;23:148695.
18. Li H, Durbin R. Inference of human population history from individual
whole-genome sequences. Nature. 2011;475:4936.
19. Xue Y, Prado-Martinez J, Sudmant PH, Narasimhan V, Ayub Q, Szpak M,
Frandsen P, et al. Mountain gorilla genomes reveal the impact of long-term
population decline and inbreeding. Science. 2015;348:2425.
20. Nadachowska-Brzyska K, Burri R, Smeds L, Ellegren H. PSMC analysis of
effective population sizes in molecular ecology and its application to black-
and-white Ficedula flycatchers. Mol Ecol. 2016;25:105872.
21. Librado P, Der Sarkssian C, Ermini L, Schubert M, Jónsson H, Albrechtsen A,
et al. Tracking the origins of Yakutian horses and the genetic basis for their
fast adaptation to subarctic environments. Proc Natl Acad Sci U S A.
2015;112:E688997.
22. Cingolani P, Platts A, Coon M, Nguyen T, Wang L, Land SJ, et al. A program
for annotating and predicting the effects of single nucleotide
polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster
strain w
1118
;iso-2; iso-3. Fly. 2012;6:8092.
23. Cingolani P, Patel VM, Coon M, Nguyen T, Land SJ, Ruden DM, et al. Using
Drosophila melanogaster as a model for genotoxic chemical mutational
studies with a new program. SnpSift Front Genet. 2012;3:35.
24. Huang DW, Sherman BT, Lempick RA. Systematic and integrative
analysis of la rge gene list using DAVID bioinformatics resources. Nat
Protoc. 2009;4:4457.
25. Benjamini Y, Hochberg Y. Controlling the false discovery rate: A practical
and powerful approach to multiple testing. J Roy Stat Soc B Met.
1995;57:289300.
26. Danecek P, Auton A, Abecasis G, Albers CA, Banks E, DePristo MA, et al. The
variant call format and VCFtools. Bioinformatics. 2011;27:21568.
27. Callicrate T, Dikow R, Thomas JW, Mullikin JC, Jarvis ED, Fleischer RC, et al.
Genomic resources for the endangered Hawaiian honeycreepers. BMC
Genomics. 2014;15:1098.
28. Tensen L, Groom RJ, Van Belkom J, Davies-Mostert HT, Marnewick K, Van
Vuuren BJ. Genetic diversity and spatial genetic structure of African wild
dogs (Lycaon pictus) in the Greater Limpopo transfrontier conservation area.
Conserv Genet. 2016;17:785. doi:10.1007/s10592-016-0821-x.
29. Kent WJ, Sugnet WC, Furey TS, Rosin KM, Pringle TH, Zahler AM, et al. The
human genome browser at UCSC. Genome Res. 2002;12:9961006.
30. Koepfli K-P, Pollinger J, Godinho R, Robinson J, Lea A, Hendricks S, et al.
Genome-wide evidence reveals that African and Eurasian jackals are distinct
species. Curr Biol. 2015;25:215865.
31. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular
Evolutionary Genetics Analysis Version 6.0. Mol Biol Evol. 2013;30:27259.
32. Katoh K, Standley DM. MAFFT multiple sequence alignment software
version 7: improvements in performance and usability. Mol Biol Evol.
2013;30:77280.
33. Price MN, Dehal PS, Arkin AP. FastTree computing large minimum
evolution trees with profiles instead of a distance matrix. Mol Biol Evol.
2009;26:164150.
34. Yang Z. PAML 4: Phylogenetic Analysis by Maximum Likelihood. Mol Biol
Evol. 2007;24:158691.
35. Finch TM, Zhao N, Korkin D, Frederick KH, Eggert LS. Evidence of positive
selection in mitochondrial complexes I and V of the African Elephant. PLoS
ONE. 2014;9, e92587.
36. Melo-Ferreira J, Vilela J, Fonseca MM, Da Fonseca RR, Boursot P, Alves PC.
The elusive nature of adaptive mitochondrial DNA evolution of an Arctic
lineage prone to frequent introgression. Genome Biol Evol. 2014;6:88696.
37. Gorman ML, Mills MG, Raath JP, Speakman JR. High hunting costs make
African wild dogs vulnerable to kleptoparasitism by hyaenas. Nature.
1998;391:47981.
38. Cadieu E, Neff MW, Quignon P, Walsh K, Chase K, Parker HG, et al. Coat
variation in the domestic dog is governed by variants in three genes.
Science. 2009;326:1503.
39. Candille SI, Kaelin CB, Cattanach BM, Yu B, Thompson DA, Nix MA, et al. A
β-defensin mutation causes black coat color in domestic dogs. Science.
2007;318:141823.
40. Clark LA, Wahl JM, Rees CA, Murphy KE. Retrotransposon insertion in SILV is
responsible for merle patterning of the domestic dog. Proc Natl Acad Sci U
S A. 2006;103:137681.
41. Kerns JA, Newton J, Berryere TG, Rubin EM, Cheng J-F, Schmutz SM, et al.
Characterization of the dog Agouti gene and a nonagouti mutation in
German Shepherd dogs. Mamm Genome. 2004;15:798808.
42. Drögemüller C, Philipp U, Haase B, Günzel-Apel A-R, Leeb T. A noncoding
melanophilin gene (MLPH) SNP at the splice donor of exon I represents a
candidate causal mutation for coat color dilution in dogs. J Heredity.
2007;98:46873.
43. Newton JM, Wilkie AL, He L, Jordan SA, Metallinos DL, Holmes NG, et al.
Melanocortin 1 receptor variation in the domestic dog. Mamm Genome.
2000;11:2430.
44. Philipp U, Hamann H, Mecklenburg L, Nishino S, Mignot E, Günzel-Apel
A-R, et al. Polymorphisms within the canine MLPH gene are associated with
dilute coat color in dogs. BMC Genet. 2005;16:34.
45. Schmutz SM, Berryere TG, Goldfinch AD. TYRP1 and MC1R genotypes and
their effects on coat color in dogs. Mamm Genome. 2002;13:3807.
46. Schmutz SM, Berryere TG, Dreger DL. MITF and white spotting in dogs: a
population study. J Heredity. 2009;100:S6674.
47. Zhang Z, Schwartz S, Wagner L, Miller W. A greedy algorithm for aligning
DNA sequences. J Comput Biol. 2000;7:20314.
48. Spiering PA, Szykman Gunther M, Wildt DE, Somers MJ, Maldonado JE.
Sampling error in non-invasive genetic analyses of an endangered social
carnivore. Conserv Genet. 2009;10:20057.
49. Spiering PA, Somers MJ, Maldonado JE, Wildt DE, Szykman GM.
Reproductive sharing and proximate factors mediating cooperative
breeding in the African wild dog (Lycaon pictus). Behav Ecol Sociobiol.
2010;64:58392.
Campana et al. BMC Genomics (2016) 17:1013 Page 9 of 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
50. Rohland N, Reich D. Cost-effective, high-throughput DNA sequencing
libraries for multiplexed target capture. Genome Res. 2012;22:93946.
51. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina
sequence data. Bioinformatics. 2014;30:211420.
52. Andrews S. FastQC: a quality control tool for high throughput sequence
data. Version 0.11.2. http://www.bioinformatics.babraham.ac.uk/projects/
fastqc/. Accessed 11 Dec 2014.
53. MagočT, Salzberg SL. FLASH: fast length adjustment of short reads to
improve genome assemblies. Bioinformatics. 2011;27:295763.
54. Li W, Godzik A. Cd-hit: a fast program for clustering and comparing large
sets of protein or nucleotide sequences. Bioinformatics. 2006;22:16589.
55. Li H, Durbin R. Fast and accurate short read alignment with Burrows-
Wheeler transform. Bioinformatics. 2009;25:175460.
56. Li H, Durbin R. Fast and accurate long-read alignment with Burrows-
Wheeler transform. Bioinformatics. 2010;26:58995.
57. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, et al. The
Sequence Alignment/Map format and SAMtools. Bioinformatics.
2009;25:20789.
58. Li H. A statistical framework for SNP calling, mutation discovery, association
mapping and population genetical parameter estimation from sequencing
data. Bioinformatics. 2011;27:298793.
59. Simão FA, Waterhouse RM, Ioannadis P, Kriventseva EV, Zdobnov M. BUSCO:
assessing genome assembly and annotation completeness with single-copy
orthologs. Bioinformatics. 2015;31:32102.
We accept pre-submission inquiries
Our selector tool helps you to find the most relevant journal
We provide round the clock customer support
Convenient online submission
Thorough peer review
Inclusion in PubMed and all major indexing services
Maximum visibility for your research
Submit your manuscript at
www.biomedcentral.com/submit
Submit your next manuscript to BioMed Central
and we will help you at every step:
Campana et al. BMC Genomics (2016) 17:1013 Page 10 of 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Genetic and morphological studies suggest that populations of L. pictus in eastern and southern Africa are distinct subspecies (Campana et al. 2016), but these have not been formally recognized. Lycaon pictus underwent two or more effective population size reductions in the last 1,000,000 years resulting in individual-specific regions of low diversity (Campana et al. 2016). ...
... Genetic and morphological studies suggest that populations of L. pictus in eastern and southern Africa are distinct subspecies (Campana et al. 2016), but these have not been formally recognized. Lycaon pictus underwent two or more effective population size reductions in the last 1,000,000 years resulting in individual-specific regions of low diversity (Campana et al. 2016). It is not certain if the low-diversity regions are characteristic of the entire population. ...
... It is not certain if the low-diversity regions are characteristic of the entire population. Campana et al. (2016) also reported that positive selection was evident on the mitochondrial genome of L. pictus, and candidate genes (i.e., agouti signaling peptide [ASIP], melanophilin [MLPH], microphthalmia-associated transcription factor [MITF], and premelanosome protein [PMEL]) might play a role in the characteristic pelage of L. pictus. ...
Article
Full-text available
Lycaon pictus (Temminck, 1820), the African wild dog, is a moderately sized carnivore with dog-like appearance and irregularly mottled black, yellow-brown, and white pelage. It has a head–body length of 76–112 cm, tail length of 30–41 cm, shoulder height of 61–78 cm, and body weight of 17–36 kg. Lycaon pictus has four toes on each foot, differentiating it from other canids; is the only extant species within the genus with no subspecies; and is unlikely to be confused with any other canid. Lycaon pictus was once widespread throughout sub-Saharan Africa inhabiting nearly all environments and now inhabits grasslands, montane savanna, and open woodlands. Lycaon pictus is recognized as “Endangered” (EN) by the U.S. Fish and Wildlife Service and the International Union for Conservation of Nature.
... Among the different canid species that went through bottlenecks or population reductions, [29][30][31][32][33][34] the Sardinian dhole showed decreased genetic diversity across the entire genome, comparable to the AHDs from Zimbabwe, Kenya, and South Africa ( Figure 4A) that went through stable and long-term population declines. 35,36 This (C) DiscoVista relative frequency analysis. Each box title indicates the corresponding branch on the tree in (A) and shows the frequency of three topologies around the focal internal branch of ASTRAL species tree. ...
... The dataset used in this study is represented by 50 representative canid genomes (Table S3), 49 of which were previously published. 17,23,35,36,40,[87][88][89][90][91][92][93][94][95][96] The genomes considered for this study were chosen to represent the genetic diversity of 9 different species (with the domestic dog considered as a different species from the gray wolf, Table S3) from Africa, Eurasia, and North-America. ...
Article
The Sardinian dhole (Cynotherium sardous) was an iconic and unique canid species that was endemic to Sardinia and Corsica until it became extinct at the end of the Late Pleistocene. Given its peculiar dental morphology, small body size, and high level of endemism, several extant canids have been proposed as possible relatives of the Sardinian dhole, including the Asian dhole and African hunting dog ancestor. Morphometric analyses have failed to clarify the evolutionary relationship with other canids.We sequenced the genome of a ca-21,100-year-old Sardinian dhole in order to understand its genomic history and clarify its phylogenetic position. We found that it represents a separate taxon from all other living canids from Eurasia, Africa, and North America, and that the Sardinian dhole lineage diverged from the Asian dhole ca 885 ka. We additionally detected historical gene flow between the Sardinian and Asian dhole lineages, which ended approximately 500-300 ka, when the land bridge between Sardinia and mainland Italy was already broken, severing their population connectivity. Our sample showed low genome-wide diversity compared to other extant canids—probably a result of the long-term isolation—that could have contributed to the subsequent extinction of the Sardinian dhole.
... Such reference genomes are particularly valuable for species of conservation concern, where accurate genomic information is crucial for effective conservation efforts and understanding the species' biology [10,13]. The significant application potential of genomics in addressing conservation problems is well-documented for various carnivorans, e.g., African wild dog Lycaon pictus, Eastern wolf Canis lupus lycaon, puma Puma concolor, Iberian lynx Lynx pardinus, and wolverine Gulo gulo [11,[30][31][32]. ...
Article
Full-text available
The European mink Mustela lutreola (Mustelidae) ranks among the most endangered mammalian species globally, experiencing a rapid and severe decline in population size, density, and distribution. Given the critical need for effective conservation strategies, understanding its genomic characteristics becomes paramount. To address this challenge, the platinum-quality, chromosome-level reference genome assembly for the European mink was successfully generated under the project of the European Mink Centre consortium. Leveraging PacBio HiFi long reads, we obtained a 2586.3 Mbp genome comprising 25 scaffolds, with an N50 length of 154.1 Mbp. Through Hi-C data, we clustered and ordered the majority of the assembly (>99.9%) into 20 chromosomal pseudomolecules, including heterosomes, ranging from 6.8 to 290.1 Mbp. The newly sequenced genome displays a GC base content of 41.9%. Additionally, we successfully assembled the complete mitochondrial genome, spanning 16.6 kbp in length. The assembly achieved a BUSCO (Benchmarking Universal Single-Copy Orthologs) completeness score of 98.2%. This high-quality reference genome serves as a valuable genomic resource for future population genomics studies concerning the European mink and related taxa. Furthermore, the newly assembled genome holds significant potential in addressing key conservation challenges faced by M. lutreola. Its applications encompass potential revision of management units, assessment of captive breeding impacts, resolution of phylogeographic questions, and facilitation of monitoring and evaluating the efficiency and effectiveness of dedicated conservation strategies for the European mink. This species serves as an example that highlights the paramount importance of prioritizing endangered species in genome sequencing projects due to the race against time, which necessitates the comprehensive exploration and characterization of their genomic resources before their populations face extinction.
... The KNP African wild dog population showed significant heterozygosity excess (P << 0.0001), suggesting that this population is not at mutation-drift equilibrium, providing strong evidence for a population bottleneck. There is supportive evidence that southern (South Africa, Namibia, Botswana, Zimbabwe, Zambia) and eastern (Mozambique, Tanzania, Kenya, Ethiopia) African wild dog populations may have also suffered from bottlenecks 20,22,47 . These results suggest that the current KNP African wild dog population was derived from a small group of individuals (resulting from a bottleneck) with limited levels of genome variation. ...
Article
Full-text available
African wild dogs (Lycaon pictus) have undergone severe population reductions and are listed as endangered on the International Union for Conservation of Nature Red List. Small, isolated populations have the potential to suffer from threats to their genetic diversity that may impact species viability and future survival. This study provides the first set of population-wide genomic data to address conservation concerns for this endangered species. Whole genome sequencing data were generated for 71 free-ranging African wild dogs from the Kruger National Park (KNP), South Africa, and used to estimate important population genomic parameters. Genomic diversity metrics revealed that variation levels were low; however, this African wild dog population showed low levels of inbreeding. Very few first- and second-order relationships were observed in this cohort, with most relationships falling into the third-order or distant category. Patterns of homozygosity could have resulted from historical inbreeding or a loss in genome variation due to a population bottleneck. Although the results suggest that this stronghold African wild dog population maintains low levels of inbreeding, likely due to their cooperative breeding system, it may lead to a continuous population decline when a reduced number of suitable mates are available. Consequently, the low genomic variation may influence species viability over time. This study highlights the importance of assessing population genomic parameters to set conservation priorities. Future studies should include the investigation of the potential of this endangered species to adapt to environmental changes considering the low genomic diversity in this population.
... In the Samburu ecosystem, illegal killing rates were strongly correlated with black market ivory prices (Wittemyer et al. 2014). Human-carnivore conflicts and retaliatory killings for livestock losses are causing declines in carnivorous species like lions, hyenas, and wild dogs (Campana et al. 2016;Carter et al. 2018;Mitchell et al. 2019;Ndeereh et al. 2019). Primate communities are mainly threatened by habitat fragmentation, loss, and degradation as well as by the decline in perennial water sources (Butynski and Jong 2014). ...
... Triangles represent DKFs and circles coyotes. (Data organized and map generated using Esri ArcMap 10.PCA of all kit fox samples (n = 70); PC1 accounts for 14.8% of variation; PC2 5.1%.Supporting InformationLycaon pictus probe designUsing two previously reported African wild dog (Lycaon pictus) genomes(Campana et al. 2016), we designed a 20,000-probe set to capture 19,729 nuclear SNPs that were polymorphic between the two genomes, along with the complete Lycaon mitochondrial genome (271 probes total). Paired raw reads from the two genomes were trimmed using Trimmomatic 0.33(Bolger et al. 2014) with the parameters ILLUMINACLIP:NexteraPE-PE.fa:2:30:10 LEADING:3 TRAILING:3 SLIDINGWINDOW:4:28 MINLEN:36. ...
Preprint
Full-text available
Genomic resources are important for evaluating genetic diversity and supporting conservation efforts. The garden dormouse ( Eliomys quercinus ) is a small rodent that has experienced one of the most severe modern population declines in Europe. We present a high-quality haplotype-resolved reference genome for the garden dormouse, and combine comprehensive short and long-read transcriptomics datasets with homology-based methods to generate a highly complete gene annotation. Demographic history analysis of the genome revealed a sharp population decline since the last interglacial, indicating that colder climates caused severe population declines prior to anthropogenic influence. Using our genome and genetic data from 100 individuals, largely sampled in a citizen-science project across the contemporary range, we conducted the first population genomic analysis for this species to investigate patterns of connectivity between regions and factors explaining population declines. We found clear evidence for population structure across the species’ core Central European range. Notably, our data provide strong evidence that the Alpine population, characterized by strong differentiation likely due to habitat isolation, represents a differentiated evolutionary significant unit (ESU). Our data also show that the predominantly declining Eastern European populations show signs of recent isolation, a pattern consistent with a range expansion from Western to Eastern Europe during the Holocene, leaving relict populations now facing local extinction. Overall, our findings suggest that garden dormouse conservation may be enhanced in Europe through designation of ESUs.
Article
Full-text available
Africa is considered the "mother continent" from which the first hominins arose. The diversified wildlife and flora of Africa, which ranges from those in the scorching Sahara and Kalahari deserts to those in the vast Savannas and wet tropical forests, are also the most diverse of any continent. Although the continent's abundance and diversity of living resources have provided critical means of subsistence for its inhabitants, future utilization of this biodiversity will demand a fundamental understanding of genetic variation and its adaptive capabilities in the face of natural and man-made stressors. Molecular ecological insights have previously been gained from a variety of vertebrate species native to Africa, and some of these discoveries have larger evolutionary and conservation implications. Despite lagging in genomics research, African scientists are increasingly eager to use the increasingly accessible -omics technology to routinely sequence more animals and plants native to Africa. This overview, which focuses on Africa's vertebrate biodiversity, aims to provide a continental scale perspective on organismal and ecological adaptations discovered through prior genomics research, as well as what conceptually these findings suggest for future research.
Article
Full-text available
Mycobacterium bovis and other Mycobacterium tuberculosis complex (MTBC) pathogens that cause domestic animal and wildlife tuberculosis have received considerably less attention than M. tuberculosis, the primary cause of human tuberculosis (TB). Human TB studies have shown that different stages of infection can exist, driven by host–pathogen interactions. This results in the emergence of heterogeneous subpopulations of mycobacteria in different phenotypic states, which range from actively replicating (AR) cells to viable but slowly or non-replicating (VBNR), viable but non-culturable (VBNC), and dormant mycobacteria. The VBNR, VBNC, and dormant subpopulations are believed to underlie latent tuberculosis (LTB) in humans; however, it is unclear if a similar phenomenon could be happening in animals. This review discusses the evidence, challenges, and knowledge gaps regarding LTB in animals, and possible host–pathogen differences in the MTBC strains M. tuberculosis and M. bovis during infection. We further consider models that might be adapted from human TB research to investigate how the different phenotypic states of bacteria could influence TB stages in animals. In addition, we explore potential host biomarkers and mycobacterial changes in the DosR regulon, transcriptional sigma factors, and resuscitation-promoting factors that may influence the development of LTB.
Article
Full-text available
Understanding predator population dynamics is important for conservation management because of the critical roles predators play within ecosystems. Noninvasive genetic sampling methods are useful for the study of predators like canids that can be difficult to capture or directly observe. Here, we introduce the FAECES* method (Fast and Accurate Enrichment of Canid Excrement for Species *and other analyses) which expands the toolbox for canid researchers and conservationists by using in-solution hybridization sequence capture to produce SNP genotypes for multiple canid species from scat-derived DNA using a single enrichment. We designed a set of hybridization probes to genotype both coyotes (Canis latrans) and kit foxes (Vulpes macrotis) at hundreds of polymorphic single nucleotide polymorphism (SNP) loci and we tested the probes on both tissues and field-collected scat samples. We enriched and genotyped by sequencing 52 coyote and 70 kit fox scats that we collected in and around a conservation easement in the Nevada Mojave Desert. We demonstrate that the FAECES* method produces genotypes capable of differentiating coyotes and kit foxes, identifying individuals and their sex, and estimating genetic diversity and effective population sizes, even using highly degraded, low-quantity DNA extracted from scat. We found that the study area harbors a large and diverse population of kit foxes and a relatively smaller population of coyotes. By replicating our methods in the future, conservationists can assess the impacts of management decisions on canid populations. The method can also be adapted and applied more broadly to enrich and sequence multiple loci from any species of interest using scat or other noninvasive genetic samples.
Article
Full-text available
DAVID bioinformatics resources consists of an integrated biological knowledgebase and analytic tools aimed at systematically extracting biological meaning from large gene/protein lists. This protocol explains how to use DAVID, a high-throughput and integrated data-mining environment, to analyze gene lists derived from high-throughput genomic experiments. The procedure first requires uploading a gene list containing any number of common gene identifiers followed by analysis using one or more text and pathway-mining tools such as gene functional classification, functional annotation chart or clustering and functional annotation table. By following this protocol, investigators are able to gain an in-depth understanding of the biological themes in lists of genes that are enriched in genome-scale studies.
Article
Full-text available
The Greater Limpopo Transfrontier Conservation Area (GLTFCA) is one of the last refuges for the endangered African wild dog and hosts roughly one-tenth of the global population. Wild dogs in this area are currently threatened by human encroachment, habitat fragmentation and scarcity of suitable connecting habitat between protected areas. We derived genetic data from mitochondrial and nuclear markers to test the following hypotheses: (i) demographic declines in wild dogs have caused a loss of genetic variation, and (ii) Zimbabwean and South African populations in the GLTFCA have diverged due to the effects of isolation and genetic drift. Genetic patterns among five populations, taken with comparisons to known information, illustrate that allelic richness and heterozygosity have been lost over time, presumably due to effects of inbreeding and genetic drift. Genetic structuring has occurred due to low dispersal rates, which was most apparent between Kruger National Park and the Zimbabwean Lowveld. Immediate strategies to improve gene flow should focus on increasing the quality of habitat corridors between reserves in the GLTFCA and securing higher wild dog survival rates in unprotected areas, with human-mediated translocation only undertaken as a last resort.
Article
Full-text available
Climatic fluctuations during the Quaternary period governed the demography of species and contributed to population differentiation and ultimately speciation. Studies of these past processes have previously been hindered by a lack of means and genetic data to model changes in effective population size (Ne) through time. However, based on diploid genome sequences of high quality, the recently developed pairwise sequentially Markovian coalescent (PSMC) can estimate trajectories of changes in Ne over considerable time periods. We applied this approach to re-sequencing data from nearly 200 genomes of four species and several populations of the Ficedula species complex of black-and-white flycatchers. Ne curves of Atlas, collared, pied and semicollared flycatcher converged 1-2 million years ago (mya) at an Ne of ≈ 200,000, likely reflecting the time when all four species last shared a common ancestor. Subsequent separate Ne trajectories are consistent with lineage splitting and speciation. All species showed evidence of population growth up until 100-200 thousand years ago (kya), followed by decline and then start of a new phase of population expansion. However, timing and amplitude of changes in Ne differed among species and for pied flycatcher the temporal dynamics of Ne differed between Spanish birds and Central/Northern European populations. This cautions against extrapolation of demographic inference between lineages and calls for adequate sampling to provide representative pictures of the coalescence process in different species or populations. We also empirically evaluate criteria for proper inference of demographic histories using PSMC and arrive at recommendations of using sequencing data with a mean genome coverage of ≥18X, a per-site filter of ≥10 reads and no more than 25% of missing data. This article is protected by copyright. All rights reserved.
Article
Full-text available
Significance Yakutia is among the coldest regions in the Northern Hemisphere, showing ∼40% of its territory above the Arctic Circle. Native horses are particularly adapted to this environment, with body sizes and thick winter coats minimizing heat loss. We sequenced complete genomes of two ancient and nine present-day Yakutian horses to elucidate their evolutionary origins. We find that the contemporary population descends from domestic livestock, likely brought by early horse-riders who settled in the region a few centuries ago. The metabolic, anatomical, and physiological adaptations of these horses therefore emerged on very short evolutionary time scales. We show the relative importance of regulatory changes in the adaptive process and identify genes independently selected in cold-adapted human populations and woolly mammoths.
Article
The common approach to the multiplicity problem calls for controlling the familywise error rate (FWER). This approach, though, has faults, and we point out a few. A different approach to problems of multiple significance testing is presented. It calls for controlling the expected proportion of falsely rejected hypotheses — the false discovery rate. This error rate is equivalent to the FWER when all hypotheses are true but is smaller otherwise. Therefore, in problems where the control of the false discovery rate rather than that of the FWER is desired, there is potential for a gain in power. A simple sequential Bonferronitype procedure is proved to control the false discovery rate for independent test statistics, and a simulation study shows that the gain in power is substantial. The use of the new procedure and the appropriateness of the criterion are illustrated with examples.
Article
The interaction between two genes, Agouti and Melanocortin-1 receptor (Mc1r), produces diverse pigment patterns in mammals by regulating the type, amount, and distribution pattern of the two pigment types found in mammalian hair: eumelanin (brown/black) and pheomelanin (yellow/red). In domestic dogs (Canis familiaris), there is a tremendous variation in coat color patterns between and within breeds; however, previous studies suggest that the molecular genetics of pigment-type switching in dogs may differ from that of other mammals. Here we report the identification and characterization of the Agouti gene from domestic dogs, predicted to encode a 131-amino-acid secreted protein 98% identical to the fox homolog, and which maps to chromosome CFA24 in a region of conserved linkage. Comparative analysis of the Doberman Pinscher Agouti cDNA, the fox cDNA, and 180 kb of Doberman Pinscher genomic DNA suggests that, as with laboratory mice, different pigment-type-switching patterns in the canine family are controlled by alternative usage of different promoters and untranslated first exons. A small survey of Labrador Retrievers, Greyhounds, Australian Shepherds, and German Shepherd Dogs did not uncover any polymorphisms, but we identified a single nucleotide variant in black German Shepherd Dogs predicted to cause an Arg-to-Cys substitution at codon 96, which is likely to account for recessive inheritance of a uniform black coat.
Article
African wild dogs (Lycaon pictus) are endangered, having disappeared from many areas where other large carnivore species have persisted. The relative vulnerability of this species has been attributed variously to its disproportionate exposure to anthropogenic threats, limitation by larger competing predators, and Allee effects caused by obligate cooperative breeding. The natural recovery of a wild dog population living on private and community land in northern Kenya provided an opportunity to investigate these potential constraints on population growth. Within a decade the population increased from near-extinction to become the 6th largest in the world. Rates and causes of mortality, and reproductive rates, were similar on community lands, where people and livestock were abundant but competing predators suppressed, and on commercial ranches, where human and livestock densities were lower but competitors more abundant. Larger packs produced larger litters, indicating a component Allee effect. However, because pack size was unrelated to population size, growth of the population was not impeded at low densities; that is, no demographic Allee effect was detectable. These results show that, despite earlier concerns, wild dogs can achieve rapid population recovery, even in a human-dominated landscape. This recovery was probably facilitated by local pastoralist traditions, which combine vigilant herding of livestock with little or no hunting of wild prey. This success might be replicated in other areas where traditional pastoralism is still practiced.