160
Molecular approaches to the conservation and management of Bengal tiger in Nepal Author Karmacharya, Dibesh Published 2017-08 Thesis Type Thesis (PhD Doctorate) School Griffith School of Environment DOI https://doi.org/10.25904/1912/2714 Copyright Statement The author owns the copyright in this thesis, unless stated otherwise. Downloaded from http://hdl.handle.net/10072/374750 Griffith Research Online https://research-repository.griffith.edu.au

Molecular approaches to the conservation

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Molecular approaches to the conservation and managementof Bengal tiger in Nepal

Author

Karmacharya, Dibesh

Published

2017-08

Thesis Type

Thesis (PhD Doctorate)

School

Griffith School of Environment

DOI

https://doi.org/10.25904/1912/2714

Copyright Statement

The author owns the copyright in this thesis, unless stated otherwise.

Downloaded from

http://hdl.handle.net/10072/374750

Griffith Research Online

https://research-repository.griffith.edu.au

Molecular approaches to the conservation and management of Bengal tiger in Nepal

A Dissertation Submitted to School of Environment

Griffith University

In Partial Fulfillment of the Requirements for the Degree of Doctor of

Philosophy

August 2017

© 2016 Dibesh Karmacharya

Supervisors

Dr. Jean Marc Hero

Dr. Jane Hughes

Author: Dibesh Karmacharya

ID: 2909328

[email protected]

SUMMARY

Bengal tiger (Panthera tigris) is an endangered species found in the lowland areas of

Nepal. Much needs to be understood about this species, including its population size,

genetic health, habitat and overall ecosystem dynamics. Although there have been efforts

to estimate tiger population numbers using camera trapping methods, the information

obtained through such efforts has been limited and there is a need to find a better method

and technology to understand wild tigers in Nepal.

Use of genetic sampling holds significant promise in execution of landscape-level

management plans as it may allow managers to measure genetic health and even gene

flow with greater certainty. Habitat connectivity and the degree of movement from one

habitat patch to another can be inferred through understanding of species behavior, but

genetic data can give us landscape level information by directly identifying patterns of

distinctiveness, endemism and the degree to which gene flow exists between

subpopulations.

Non-invasive genetic studies on populations, such as those that use scat samples, have

increased in recent years, and have been used to estimate population size of many elusive

and endangered species. I have used this technique to understand tiger and its habitat.

In my study I have uncovered prevalent sympatric biodiversity in the tiger habitat which

aids in understanding the species conservation from an ecological perspective. Out of total

collected scat samples (n=420), only 56% (n=237) were of tiger. The remaining non-tiger

samples (n=183) included non-focal carnivores; leopard (Panthera pardus, n=83), leopard

cat (Prionailurus bengalensis, n=10), jungle cat (Felis chaus, n=2), fishing cat

(Prionailurus viverrinus, n=2) and fox (Vulpes spp., n=10). The spatial distribution of

three small felids: leopard cat, jungle cat and fishing cat in sub-tropical deciduous forest of

Terai has provided insights on sympatric occurrences of small cats with each other and

with two other large felids, leopard and tiger, in CNP.

Overall landscape level of information on tigers on such important aspects like genetic

health, sub-population structure and gene flow will help in designing broader conservation

strategies for the species. Of the 770 scat samples collected opportunistically from four

protected areas and five presumed corridors, 412 were tiger (57%). Using eight

microsatellite markers, I identified 78 individual tigers. I used this dataset to examine

population structure, genetic variation, contemporary gene flow, and potential population

bottlenecks of tigers in Nepal. I detected three genetic clusters consistent with three

demographic sub-populations and found moderate levels of genetic variation (He = 0.61,

AR = 3.51) and genetic differentiation (FST = 0.14) across the landscape. I detected 3-7

migrants, confirming the potential for dispersal-mediated gene flow across the landscape. I

found evidence of a bottleneck signature likely caused by large-scale land-use change

documented in the last two centuries in the Terai forest. Securing tiger habitat including

functional forest corridors is essential to enhance gene flow across the landscape and

ensure long-term tiger survival.

The molecular forensic tools that we have developed have been helping the law

enforcement officials in Nepal in the fight against wildlife poaching. I created Nepal‘s first

comprehensive reference genetic database of wild tigers through the Nepal Tiger Genome

Project (2011-2013). This database has nuclear DNA microsatellite genotype and sex

profiles, including geo-spatial information, of over 60% (n=120) of the wild tigers of

Nepal. I analyzed 15 cases of confiscated poached tiger parts. Ten were identified as males

and five were females, and I determined probable geo-source location for 14 of those

samples by combining inferences from two different analytical methods of population

genetic structure and phylogenetic analysis. Conclusively, eleven of the fourteen samples

were assigned to the western region of Nepal, while the remaining three were determined

to be distantly related to the others and were identified as outliers. Among these, one

sample was an exact match to a female tiger in our reference database previously detected

in Bardia National Park. My study revealed the western region, particularly Bardia, is a

poaching hotspot for illegal tiger trade across the Tera Arc Landscape in Nepal. I present

scientific evidence to incriminate criminals in a court of law and advocate for the increased

use of molecular forensics in wildlife conservation efforts.

And finally, I have mapped the inter-relationship between tiger and its gut microbiome.

This study is the first of its kind done on wild tigers; and our results show an association

between the species with its gut microbiota and habitat. The scat microbiome of the Bengal

tiger contained the phyla Proteobacteria (~37%), Firmicutes (~27.7%), Bacteroidetes

(~18.6%), Fusobacteria (~13.4%) and Actinobacteria (~2.7%), but in varying proportions

across the habitats. Predictive metagenome functional content analysis (PICRUST) of the

gut microbiome, restricted to three significantly different bacterial genera (Comomonas,

Collinsella and Fusobacteria), identified 36 significantly (p value < 0.05) differential

functional content systems. Chitwan samples contained a lower proportion of sequences

associated with immune system and infectious diseases, but higher proportion that are

associated with xenobiotic biodegradation /metabolism, cellular processing and signaling,

and neurodegenerative diseases compared to the other two. These findings underscore the

importance of a conservation design that specifically seeks to maximize genetic and gut

microbial diversity amongst wild tigers in Nepal.

My Study, through the Nepal Tiger Genome Project, has looked into a range of molecular

approaches to comprehend the tiger population from conservation genetics perspective and

uncovered wealth of information which is now being utilized to formulate effective

conservation strategies. Most of this work was done for the first time in Nepal; and some

work, especially molecular forensics and gut microbiome study, were done for the first

time. My utility based study has also been able to create necessary capacity to do similar

work in other species in developing countries like Nepal, and beyond this I am looking at

creating viral and dietary profiles on some of these tiger fecal samples as part of my

commitment to understand this species further

ACKNOWLEDGEMENTS

To my Professors Jean Marc Hero and Jane Hughes for encouraging and guiding me while

I designed and carried out my research; and a special thank you to Professor Hero for

encouraging me to pursue my PhD.

The whole study would not have been possible if it were not for the funding support from

USAID to initiate Nepal Tiger Genome Project. Special thanks to USAID/Nepal mission

director, Dr. Kevin Rushing, for believing in my tiger project and helping us with the

funding. I would like to thank all the team members who worked extremely hard in making

the project a grand success. I was privileged to lead the project as the Principal

Investigator.

A special thank you also goes to the Department of National Parks and Wildlife

Conservation and the Ministry of Forests and Soil Conservation of the Government of

Nepal for providing us with all the needed guidance and permits.

Dr. Momchilo Vuyisich of the Los Alamos Laboratory has been one of my key

collaborators who made tiger gut microbiome study possible. I am very thankful to him and

his team for providing me with the valuable support. Also a big thank you to Dr. Lisette

Waits of the University of Idaho for providing me with continued technical support on

conservation genetics.

CMDN Bioinformatics and Computational Biology team worked really hard to analyze my

metagenomics data and helped build the necessary computational pipeline. Thank you for

this wonderful team. And a big thank you to all the members of CMDN and INPL team for

helping me generate all these wonderful and important data.

And finally a wonderful appreciation to my parents and Ms. Manisha Bista for encouraging

and pushing me to get my PhD. Their constant reminder and pushing made me realize the

importance of getting this degree. A warm thank you to them

Statement of Originality

This work has not previously been submitted for a degree or diploma in any university. To

the best of my knowledge and belief, the thesis contains no material previously published

or written by another person except where due reference is made in the thesis itself.

(Signed)

Name of Student DIBESH BIKRAM KARMACHARYA

CHAPTER CONTRIBUTIONS

This work has been entirely designed, implemented and analyzed by the Center for

Molecular Dynamics Nepal (CMDN) under my leadership. Some parts of the work have

been done in collaboration with various experts. Entire work has been a part of Nepal

Tiger Genome Project which was funded by USAID (www.ntgp.org.np). I am the

principal investigator of the project and I was responsible for designing and securing the

funding. I also acted as a supervisor and lead the entire field and lab team, and fostered

collaborative relationships with the government as well as non-government agencies.

Please see table below for my contribution to conduct this research.

Chapter Title Concept Funding acquisition

Study Design

Overall Supervision

Field Work

Lab Work

Data Analysis

Manuscript Preparation

Publication Status

III

Incidental discovery of non-

focal carnivore species during

genetic study of Bengal tiger

(Panthera tigris tigris) and

snow leopard (Panthera uncia)

in Nepal.

100% 100% 100% 100% 60% 60% 80% 80% Published

Asian Journal of Conservation:

December 2016

IV Assessment of genetic diversity,

population structure, and gene

flow of tigers (Panthera tigris

tigris) across Nepal's Terai Arc

Landscape.

100% 100% 80% 100% 60% 60% 80% 70% Accepted PLOS One: Feb 2018

V Species, sex and geo-location

identification of seized tiger

(Panthera tigris tigris) parts in

Nepal- a molecular forensic

approach

100% 100% 100% 100% 60% 70% 80% 80% Review received and revised

version submitted to PLOS One: Jan

2018

VI

Gut microbio me and

their putative metabolic

functions in fragmented

Bengal t iger population

of Nepal

100% 100% 100% 100% 100% 70% 80% 90% Manuscript Submitted to

Nature: Jan 2018

Contents

SUMMARY

ACKNOWLEDGEMENTS

Statement of Originality

CHAPTER 1

1. INTRODUCTION

CHAPTER II

2. METHODOLOGY & STUDY DESIGN

3. CHAPTER III

Incidental discovery of nonfocal carnivore species during genetic study of Bengal tiger

(Panthera tigris tigris) and snow leopard (Panthera uncia) in Nepal

4. CHAPTER IV

Assessment of Genetic Diversity, Genetic Structure, and Gene Flow of tigers (Panthera tigris

tigris) in the Terai Arc Landscape of Nepal

5. CHAPTER V

A molecular forensic approach for species, sex and geosource location identification of

seized tiger (Panthera tigris tigris) parts in Nepal

6. CHAPTER VI

Gut microbio me and t heir putative metabol ic functions in fragmented B engal

t iger populat ion of Nepal

7. Chapter VII

Conclusion and Future Direction

CHAPTER I

INTRODUCTION

Bengal tiger (Panthera tigris) is an endangered species found in the lowland areas of Nepal. Much needs to be

understood about this species, including its population size, genetic health, habitat and overall ecosystem dynamics.

Although there have been efforts to estimate tiger population numbers using camera trapping methods, the

information obtained through such efforts has been limited and there is a need to find a better method and technology

to understand wild tigers in Nepal.

1.1 Landscape level tiger population management and conservation

There is growing emphasis on management of tigers at a landscape level in order to facilitate long-term tiger recovery

(Sanderson et. al. 2006), a practice currently seen in the Terai Arc Landscape (TAL) in India and lowland Nepal.

With one of the highest densities of wild tigers on earth, this region has become a global priority for conservation of

wild tigers (Barlow et al. 2009, Sanderson et al. 2006).

The TAL is a tiger inhabited landscape on the southern slope of the Himalayas, with highly productive alluvial

grasslands and riverine forests capable of supporting some of the highest tiger densities across the range. In this

human-dominated landscape, the Government of Nepal and partners have set an ambitious goal of sustaining 550

tigers. Demographically, core tiger populations are currently limited to three sub-populations in four protected areas

across the landscape: Chitwan population (Parsa Wildlife Reserve & Chitwan National Park), Bardia population

(Bardia National Park) and Suklaphanta population (Suklaphanta Wildlife Reserve) with an estimated total of 125

adult tigers across these areas (DNPWC 2010).

1.2 Understanding biodiversity, species presence/absence and distribution-holistic

approach to conservation

Amid concerns over the loss of biodiversity, experts have been increasingly seeking clues to understand ecosystems

and biodiversity (Cardinale et. al, 2012). The prevailing conservation and management practices, globally, include

coarse-filter and fine-filter approaches (Noss et. al., 1987).The former aims to preserve entire communities of plants

and animals by protecting large areas of habitat (ecosystem approach) whilst the latter focuses to protect species

(species approach).

Current burgeoning conservation needs and limited availability of resources often forces conservation managers and

planners to rely on information on the occurrence of apex species such as tiger in the TAL (Forrest et. al., 2012) to

serve as ecosystem health indicators (Simberloff et. al, 1998).

Core tiger populations are fragmented and found mainly in lowland areas across the four protected areas of Parsa,

Chitwan, Bardia and Shuklaphanta and occupy 36% of 26,000 sq. km of the forested landscape of TAL, Nepal

(Shrestha, 1997) (Fig. 1). This habitat is also shared with leopard (Panthera pardus), dhole (Cuon alpinus), sloth bear

(Melursus ursinus) and striped hyena (Hyaena hyaena) (Odden et. al, 2010).

In the past, there have been numerous studies conducted on the status of the flagship species (Subedi et. al, 2013).

However, there has been limited information available on the degree to which the non-focal species have benefited

while working with these flagship species. Use of the non-invasive sampling for genetic studies utilizing molecular

scatology techniques (Kelly et. al, 2012) primarily on flagship species could be beneficial in identifying the

occurrence of the non-focal species along with the species of interest, thereby helping to map and document the

spatial distribution of species.

Fig 1: Protected areas and National Parks of Nepal

1.3 Gene flow, genetic diversity and genetic structure of tiger conservation implications

Low genetic diversity and reduction in habitat area are potential influential factors leading to the extirpation of

populations and extinction of species (Lacy et. al. 1997). In the face of habitat fragmentation and isolation,

maintaining the optimum level of genetic connectivity among fragmented patches is a challenging task (Mills et. al,

1998).Yet gene flow among isolated populations is necessary to avert the negative consequences of genetic drift

and/or inbreeding and plays an important role in persistence of natural populations. Gene flow is particularly critical

for carnivores that often occur at naturally low densities thus increasing their risk of extinction due to a greater

susceptibility to stochastic events (Caughley et. al., 1994; Frankham et. al., 2002).

A recent study has shown that tigers occupy 36% of the TAL (Nepal) clustered within the protected areas (Barber-

Meyer et. al., 2013). Few tiger signs have been detected outside of protected areas (Wikramanayake et. al., 2010),

although telemetry studies have shown that tigers in the TAL have dispersed as far as 30 km (Smith et. al., 1993),

including through human dominated areas within the landscape (Sharma et. al, 2009). Camera trap data have also

confirmed the presence of tigers in the corridors (Wikramanayake et. al., 2010).Thus, a landscape level genetic

approach to conservation is needed to assess whether dispersal and subsequent breeding (genetic migration) of tiger

occurs across the fragmented habitat. Existing methods (e.g. camera trapping and telemetric studies) require long-

term data sets to study the dispersal events (Patil et. al., 2011) and are somewhat ineffective for determining numbers

of genetic migrants. Noninvasive genetic techniques, therefore, are useful tools to study landscape connectivity and

dispersal across the TAL (Joshi et. al., 2013).

1.4 DNA molecular forensics- tool for tiger anti-poaching and fight against wildlife

crime

Applications of DNA profiling in wildlife forensics remain rare, although they have been used successfully in

investigations in Canada with mule deer, white-tailed deer, moose, caribou and black bear (Ogden et al. 2009) and in

Italy in one particular poaching case of a wild boar (Lorenzini 2005) and six Italian wolves (Caniglia et al. (2009).

Use of this method in tiger conservation has also been rare, although one study demonstrated feasibility in a tiger

bone smuggling case in China (Xu et al. 2005) and in the illegal sale of tiger meat from a circus tiger (Wan and Fang

2003). Genetic tools can also provide opportunities to discern geographic origin of confiscated wildlife parts and their

derivatives (Ogden et al. 2009, Palsboll et al. 2006), information which can inform management of areas that require

greater investments in anti-poaching.

The Nepal government‘s plan also emphasizes the need for a new approach to prevent poaching and illegal trade in

tiger parts. The Nepal government has asserted that the volume of seized tiger parts is high, with Nepal acting as both

a source and as a conduit for trafficking into China/Tibet and other areas (Bajimaya et al. 2007). Current high levels

of poaching activity severely undermine protection efforts. Moreover, anti-poaching operations are highly localized

and in need of landscape-level coordination. The government of Nepal has proposed establishing a system to better

track records of poaching incidences and confiscation and enforcement of CITES by increasing capacity for

identifying animal parts and their derivatives, by using effective and modern forensic procedures. In reviewing

success of investments in tiger conservation, best practices in anti-poaching projects have been those incorporating

scientifically sound wildlife monitoring; and projects that have been unsuccessful had failed to effectively

disseminate data to management and other conservation institutions (Gratwicke et al. 2007). There is a significant

need for studies on tigers; and projects related to tiger conservation need to make reliable information readily

available to decision-makers and other stakeholders. Without accurate information on populations it is impossible to

determine if the large investments currently being made have been effective.

1.5 Profiling gut microbiome of tiger- its implication on wildlife disease dynamics and

comparative genomics

Recent studies have clearly demonstrated the enormous pathogen diversity that exists among wild animals. This

exemplifies the required expansion of our knowledge of the pathogen and gut microbiome diversity present in

wildlife (Bodewes et al. 2014). There is very limited information on prevalent diseases in wild tigers, although it is

widely believed that viral infections such as Rabies and Canine Distemper are commonly seen in captive populations.

Often, disease study is almost impossible in the wild due to difficulty in accessing samples. My study uses DNA

extracted from scat to perform metagenomic analysis using next generation DNA sequencing technology to profile

prevalent pathogens (bacteria) in the identified tigers, thereby producing a tiger gut microbiome profile for the first

time. This should provide some valuable information on pathogens present in the wild tiger population, and also by

conducting phylogenetic analysis between host (tiger) and pathogen (present bacteria), my research might be able to

create comparative gene flow information, thereby providing a new tool to estimate recent host sub-population

interactions utilizing molecular epidemiology approach. Currently, no such method is available.

1.6 Available tools for tiger research

1.6.1 Conventional methods of collecting data- Pugmark analysis

For tigers, use of pugmarks as a technique to

identify individuals and generate population estimates

has long been used by the Indian government; however,

the method‘s high susceptibility to errors generates

limited confidence in results (Karanth et al. 2003,

Narain et al. 2005). This technique relies on

measurement of paw prints of animals, with the

presumption that each paw print profile is unique to an

individual animal. However, environmental exposure

(ground texture, moisture etc) greatly influences the

paw imprints and hence there is a wide error rate on

identifying pugmark impressions.

1.6.2 Camera trapping

A recent alternative has been to use remote cameras to

photograph (camera trapping) animals under the assumption

that individuals can be identified based on certain physical

characteristics such as stripes or other markings. Rates of

appearance (detection histories) of individuals over a

survey period can be combined with capture-mark-

recapture (CMR) methods to generate robust estimates of

population size. CMR methods and analyses have a long

history and thorough vetting in wildlife management, but

only recently have newer data collection techniques such as

camera trapping been combined with well-established CMR

population estimation to obtain better, more defensible and robust estimates of population size (Karanth et

al. 2006, Mondol et al. 2009). There are however, a number of potential disadvantages of using camera

traps, including poor performance in areas of difficult terrain or adverse weather, susceptibility to theft or

vandalism, low detection and demanding field logistical support and possible violation of population

closure assumptions in mark-recapture (Mondol et al. 2006, Harihar et al. 2009). Moreover, due to

limitations of its use in areas of low population density, it may not be feasible in marginal habitats, reducing

the extent to which it can be applied in studying landscape-level population dynamics (Bhagavatula and

Singh 2006). The use of camera traps is also inherently restricted to animals that can be identified down to

the individual, due to their distinctly different coat patterns with a high degree of confidence through

photographs. Finally, threat of vandalism, potentially makes it infeasible to use camera traps in areas

outside of protected areas, such as in corridors surrounded by villages with high human use.

1.6.3 DNA technology and its utility in conservation science

A research tool yet to be fully utilized for tigers is genetic sampling in combination with Capture-Mark-Recapture

(CMR) statistical techniques; this constitutes use of molecular techniques to derive unique genetic profiles for

individuals based on collection of samples that harbor DNA such as blood, tissue, hair, bones and scats (Waits and

Paetkau 2005, Wetton at al. 2004, Ruttledge et al 2009, Gillett et al 2010, Wilson et al, 2009). Enthusiasm has

increased as studies have documented the potential use of genetic

sampling in applications such as non-invasive population estimation and

wildlife forensics (Mills et al. 2000; Kelly et al 2012). This is of

particular interest in tiger conservation, providing new opportunities for

understanding not only population size through genetic identification of

individuals, but it also gives information on landscape level population

dynamics (Holdregger and Wagner 2008, Sharma et al. 2009); and is

useful for enforcing anti-poaching laws (Singh et al. 2004, Wetton et al.

2004). Use of genetic sampling holds significant promise in execution of

landscape-level management plans as it may allow managers to measure

genetic health and even gene flow with greater certainty. Habitat

connectivity and the degree of movement from one habitat patch to

another can be inferred through understanding of species behavior, but

genetic data can give us landscape level information by directly identifying patterns of distinctiveness, endemism and

the degree to which gene flow exists between subpopulations (Cracraft et al. 1998, Holdregger and Wagner 2008).

1.6.4 Advantages of non-invasive sampling and utilization of genetic tool(s) in tiger

research

It is difficult to enumerate population of wide-ranging, elusive species that occur at low density, like tigers; non-

invasive methods of detecting tigers have been frequently employed to estimate number of individuals in a certain

area (Schipper et al., 2008). Non-invasive methods have also been favored for eliminating the need for direct

interactions that could potentially have adverse effects on animal welfare (Lukacs and Burnham 2005; Kelly et al.

2012).

Non-invasive genetics has a number of advantages in

population studies using CMR techniques, including

increasing number of observations (i.e. scats collected)

relative to camera trapping (photos collected). This

increases reliability and precision in estimates and allows

for a shorter sampling period, constituting a greater

adherence to population closure assumption in capture-

recapture (Mondol et al. 2009).

Non-invasive genetic techniques have a particular

advantage over the camera-trap method as it can be employed in areas where use of cameras may be deemed

infeasible. In addition, it can identify unique individuals from other species, which cannot be discerned from

photographs of animals without distinct coat patterns. However, although this method has a number of advantages

over other methods, the most effective use of DNA identification in population studies may be in circumstances

where it is used to compliment other methods, significantly increasing reliability in estimations (Bhagavatula and

Singh 2006). Non- invasive genetic studies on populations, such as those that use scat samples, have increased in

recent years, and have been used to estimate population size in coyotes, redwolves, elephants, European badger,

Brown bears and Scandinavian wolverines (Miller et al. 2005) and in documenting dispersal of individuals in a

population (Forbes and Boyd 1997, Janecka et al. 2008b, Onorato et al. 2009). Use of scat samples to estimate

population size of tigers, however, has been virtually non-existent despite its potential feasibility, and what work has

been done is in its infancy, occurring almost exclusively in India (Bhagavatula and Singh 2006, Narain et al. 2005).

1.7 Study Species

Bengal Tiger (Panthera tigris tigris)

The Bengal tiger (Pathera tigris tigris; Class-

Mammalia, Order-Chordata, Family- Felidae) is the

most numerous tiger subspecies found in the Indian

sub-continent (Bangladesh, Bhutan, Nepal and India)

and with the total estimated population (2011) of less

than 2,500 individuals it has been classified as

endangered by the IUCN. India has the highest

number of tigers (1,706–1,909) (Jhala, Y.V. et. al, 2008), followed by Bangladesh (around 440) (Khan, M., 2012),

Nepal (163–253) (Dhakal, M., 2014) and Bhutan (103) (WWF-―Bhutan‘s tigers, 2015).

General Body Morphology:

The Bengal tiger's has a yellow to light orange coat with stripes ranging from dark brown to black; belly

and interior parts of limbs are white, and tail is orange with black rings. Male Bengal tigers have an average

total body length (including tail) of 270-310 cm (110-120 in); while females measure 240-265 cm (94-104

in) in length on average. The average height of tigers is 90-110 cm (35-43 in) and the average body weight

ranges from 180-258 kg (397-569 lb) for males and 100-160 kg (220-350 lb) for females (Chundawat et. al.,

2015; Mazak, V., 1981).

Habitat:

In the Indian subcontinent, tigers inhabit tropical moist

evergreen forests, dry forests, tropical and subtropical

moist deciduous forests, mangroves, subtropical and

temperate upland forests, and alluvial grasslands. Their

habitat also consists of grassland and riverine and moist

semi-deciduous forests along the major river system

(Wikramanayake,1999). The tiger population in the Terai

of Nepal is split into three isolated subpopulations that are

separated by cultivation and densely settled habitat. The

largest population lives in the Chitwan National Park

(CNP) and in the adjacent Parsa Wildlife Reserve (PWR)

encompassing an area of2,543 km2 (982 sq mi) of prime

lowland forest. To the west, the Chitwan population is

isolated from the one in Bardia National Park (BNP) and adjacent unprotected habitat farther west, extending to

within 15 km (9.3 mi) of the Shuklaphanta Wildlife Reserve (SWR), which harbors the smallest population. The

bottleneck between the Chitwan-Parsa and Bardia-Sukla Phanta metapopulations is situated just north of the town of

Butwal (Smith, J. et. al., 1998).

As of 2009, an estimated 121 breeding tigers lived in Nepal. By 2010, the number of adult tigers had reached 155.A

survey conducted from December 2009 to March 2010 indicates that 125 adult tigers live in CNP and its border areas

covering 1,261 km2 (487 sq mi) (WWF, 2009). From February to June 2013, a camera trapping survey was carried

out in the TAL, covering an area of 4,841 km2 (1,869 sq mi) in 14 districts. The country‘s tiger population was

estimated at 163–235 breeding adults comprising 102–152 tigers in the Chitwan-Parsa protected areas, 48–62 in the

Bardia-Banke National Parks and 13–21 in the SWR (Dhakal, M., 2014). Under the project- Nepal Tiger Genome

Project (2011-2013), Nepal carried out the first comprehensive genetic study and survey of tigers and estimated

overall population to be around 220 in 3 major national parks and connecting (and functioning) corridors

(Karmacharya et. al., 2014-unpublished work).

Conservation Threats:

Over the past century the tiger population has dwindled and its habitat reduced drastically. None of the tiger

conservation landscapes is large enough to support an effective population size of 250 individuals. Habitat loss and

high incidences of poaching are serious threats to tiger‘s survival (Chundawat, et.al., 2011). Additionally, emerging

and re-emerging diseases might be the next big threat to tiger conservation.

Poaching

Immediate threat to wild tiger populations in south Asia is poaching. There is often inadequate enforcement response

in these countries to stop or slow down wildlife crime (poaching). Various tiger body parts such as skin and bones

have high monetary value in countries like China where they are used as ornamental and decorative items and as

ingredients in traditional medicine (Hemley, et. al., 1999). According to the report from Wildlife Protection Society

of India, in 15 years (1994-2009) there were more than 893 documented cases of tiger killings in India, which

experts believe is just a fraction of actual total poaching.

Human-Wildlife conflict

The increasing human population is encroaching and shrinking tiger habitats thereby increasing the risks of human-

tiger conflict. The conflict often results in human, tiger or livestock losses. There were no recorded cases of human

death due to tiger attacks before 1980 in the CNP of Nepal. However the following year 13 people were killed in tiger

attacks (McDougal, 1987). Every year there are reported cases of human and livestock casualties in Nepal, and the

government is struggling with mitigation and compensation challenges with human wildlife conflicts. This creates

resentment in the affected communities, and often this result in retaliatory killing of animals.

1.8 Study Aims and Hypothesis:

Conservation of the Bengal tigers in Nepal is a top priority due to the species‘ dwindling numbers worldwide. It is

also important to have a comprehensive understanding of tiger conservation from a landscape perspective. In order to

develop effective policies and strategies for the conservation of Bengal tigers, effective census tools need to be

developed and deployed. The previous traditional/conventional tools for conservation are often inadequate and

ineffective to address the problem.

Genetic analysis has become an effective and popular method and is used in all aspects of wildlife biology and

conservation. Since portions of the genome of every individual are unique, the use of genetic tools yield highly

specific information that can help in estimating animal migration rates, population size, bottlenecks and kinship.

Genetic analysis can also be utilized to identify species, sex and individuals; and provide insight on its population

trend as well as to gather other taxonomic level information. Since it is infeasible to enumerate populations of low

density, wide-ranging and elusive species like tigers, non-invasive methods of detecting the species by using scat or

fecal sample have been frequently employed to infer estimations on the number of individuals in a certain area;

moreover, this method has also been favored for eliminating the need for direct interactions (invasive) that could

potentially have adverse effects on animal welfare. Scat is also a good source of DNA of gut microbiome of tigers.

The implication and utility of information obtained through this research has great potential to address many

outstanding tiger related conservation questions- How many tigers are there in each of our national parks? How

healthy are they from genetic perspective? Are there any interactions between pockets of tiger populations? Will we

be able to track the source of poached tiger parts and help in the fight against wildlife crime? These are all important

questions that can only be answered through genetics and use of DNA technology. Additionally, by building a

genetic database and creating a geo-spatial map of all the wild tigers in Nepal, one could potentially track each tiger

over time and gauge its status.

The next big threat to tiger conservation will be coming from wildlife diseases; as the tiger habitats are gradually

being encroached and circled by humans and livestock, there is a growing threat of zoonotic diseases being

transmitted between humans, animals (livestock) and wildlife populations. We know very little about the gut

microbiome of wild tigers and its role in general health and disease implications on the species. My research will shed

new light on gut microbiota of wild tigers in Nepal. This research and its findings will pave ways to conduct similar

research on other endangered and important species of Nepal.

My research thesis has 6 main chapters, including methodology and study design data (Chapter II) that describes the

overall method that was used to derive all the experiments. The first research chapter (Chapter III) outlines discovery

of other sympatric carnivore species while we conducted the tiger study- the findings are important to understand

conservation of the species at a landscape scale with holistic approach. In the second data chapter (Chapter IV),

tiger‘s genetic structure, gene flow and population genetics analysis is described in detail. The practical utility of tiger

genetic database created as a result of our project (Nepal Tiger Genome Project) has been providing valuable

information for identification of at source location of poached tiger parts; this molecular forensic tool has now been

serving Nepal Police in its wildlife crime investigations and this has been described in Chapter V. The gut

mircrobiome of wild tigers, and comparative analysis between host (tiger) and its gut microbiome (including potential

pathogens) is analyzed and described in the final chapter (Chapter VI).

Fig 1: Sample collection area- total of 108 grids across Terai

Arc. Each grid covering an area of 225 sq km

CHAPTER II

METHODOLOGY & STUDY DESIGN:

Field sample collection, laboratory processing and data analysis relevant to more than one chapter are described here.

Methods and analysis pertaining to specific questions are provided in the relevant chapters. Overall my research

consists of the following methodology-

2.1Field Method for Collection of tiger Scats The basic conceptual framework was to apply a

grid across the TAL and systematically search

for tiger scats. The landscape was divided into a

grid-cell of size 15 by 15 km. This sampling

unit of 225 km2 corresponds to the home range

size of a male tiger, and will yield a total of 108

grids across the landscape. The national parks

have a network of the fire lines, trails, elephant

routes, and game-trails, all well distributed

spatially, providing potential sites for detecting

scats. Within each grid cell, spatially distributed

routes were identified by searching on foot. To

remain consistent with previous surveys, effortwas

made to match previous (2009) occupancy survey for new scat searches.

Field sample (scat) collection and preservation;

Genetic analysis- Species, Sex and Individual Identification; meta-genomics

analysis for bacterial profiling

Statistical and Bioinformatics Analyses-Use of analytical tools for data analysis

Building Geo-Spatial database

Scats were collected from the three sub-populations (CNP – PWR, BNP, and SWR) where likelihood of encountering

tiger scats was fairly high. To maximize time and resource efficiency; high detection probability and occupancy (Psi)

value grids from the previous survey were searched first. Scats were collected from grids across six critical sites

identified in the landscape. The critical sites included functional corridors (Khata, Basanta, Laljhad) and dispersal

bottlenecks (Mahadevpuri, Lamahi and Dovan). During the second and last phase, we used village rangers

(Bhagheralu) who have good knowledge about where wild animals (including tigers) roam in and around forest.

There is an existing network of village rangers across the Terai Arc.

Fig 2: Known Site Occurrence (Psi) of Tiger across Terai Arc Landscape where we will target scat collection.

[Note: Sample size in each chapter takes specific samples for study consideration. Since this is fairly large project, not

all studies incorporated same samples. Same is true with the number of samples that have been considered for

certain application, hence sample numbers in each chapter vary. Above described sampling grids are aligned with the

camera trap survey for comparative analysis between camera trap and genetic based population census. The

population census and estimation is not included in this Dissertation Report.]

2.2Field Methods for Preservation of Scats

Upon encounter, each scat was categorized by age, based on color, moisture, odor strength, and presence and absence

of mould. Vegetation measurements, weather and temperature at the site were recorded. Additionally, time and

location were also recorded on a handheld GPS unit.

Scrapings from the outer surface of each scat were collected. Collected scat samples were stored in 2-ml vials in 1:4

ratio with DET buffer. Each vial containing DET buffer contains 20% Di-methyl sulfoxide DMSO), 0.25

Methylenediamene tetra-acetic acid (EDTA), 100 mMTris, pH 7.5 and saturated with NaCl.

2.3 Field Safety & Minimizing environmental impacts from the research

Standard Operating Protocol (SOP) was developed incorporating all the safety training for field personnel. Trainings

were provided to the field personnel prior to their deployment on sample identification, sample collection & storage,

field safety and other environmental considerations. The study was based on non-invasive samples (only fecal

matter); we did not disturb or harm any animals. The technology allowed us to extract DNA from fecal matter to

obtain necessary information.

2.4Genetic Analysis:

Some portion of genetic analysis of tiger scats was adapted from earlier work by Bhagavatula et al. (2009) from the

Centre for Cellular and Molecular Biology, India (Genotyping faecal samples of Bengal tiger Panthera tigris tigris

for population estimation: A pilot study. BMC Genetics. 2006).

2.4.1Sample processing and DNA extraction from scat:

DNA from scat samples was extracted using the QIAamp DNA Stool kit

(Qiagen, Germany).A pea-sized sample was scraped from the outer surface

of scat with sterile tweezers and scissors in a weigh boat with 2ml of ASL

buffer. After making a suspension by crushing with tweezers, it was

transferred to a 2 ml tube. Genomic DNA was extracted by using the Kit‘s

protocol.

2.4.2Tiger-Species Identification:

Species identification was done by tiger specific target amplification in the

mitochondrial cytochrome b gene producing 162bp PCR product.

The following PCR primer set was used for tiger species id PCR:

TIF: 5'-ATAAAAAATCAGGAATGGTG-3'

TIR: 5'-TGGCGGGGATGTAGTTATCA-3'

Confirmation of Tiger specific species identification was done by running these samples in general carnivore specific

PCR by using following PCR primer sets:

CYTB-SCT-F 5'-AAACTGCAGCCCCTCAGAATGATATTTGTCCTCA-3'

CYTB-SCT-R 5'-TATTCTTTATCTGCCTATACATRCACG-3'

All tiger species PCR positive samples were carnivore PCR positive.

Fig 3: Species identification PCR target.

2.4.3 Sex Identification:

Sex identification PCR primers for felids target the amelogenin region based on sequence data available for domestic

cat (GenBank accession AF114709, AF197967). In domestic cat, the Y chromosome copy (AMELY) has a 20 bp

deletion when compared to the X chromosome gene (AMELX).

PCR primers used for sex identification PCR:

AMEL-F 5'-CGAGGTAATTTTTCTGTTTACT-3'

AMEL-R 5'-GAAACTGAGTCAGAGAGGC-3'

The expected PCR banding pattern in gel electrophoresis:

Fig.4: Sex identification of tiger species positive samples using Amelogenin

gene PCR assay

2.4.4Individual Identification (DNA fingerprinting):

Tiger species PCR positive samples were further processed to complete individual identification (profiling) using

microsatellite analysis. 10 microsatellite loci were examined:

Final microsatellite markers used for individual identification:

Table1: Loci used in final analysis (with allele variation observed in the Chitwan Population)

Pop Locus No. Of alleles Allele Range

CNP

FCA205 5 100-110

FCA391 4 144-156

PttD5 2 199-207

FCA232 4 95-107

FCA304 5 121-139

F85 7 149-179

FCA043 5 113-129

F53 7 157-180

FCA441 4 107-119

PttA2 1 186

2.4.5 Statistical Analyses:

Genetic variation was assessed using MICROSAT software (Park 2001) for allele size, allelic richness, and expected

and observed heterozygosities. Tests for Hardy-Weinberg equilibrium and linkage disequilibrium was carried using

the GENEPOP software (Raymond and Rousset 1995). STRUCTURE (Pritchard et al. 2000)was used to perform a

hierarchical Bayesian clustering method, combining mtDNA and microsatellite data to yield population clusters based

on individual genotypes. Assignment tests werealso carried out in GENECLASS (Piry et al. 2004) to identify recent

dispersers and their first and second generation progeny (Rannala and Mountain 1997). Population structure was

assessed using the analysis of molecular variance (AMOVA) method in the program ARLEQUIN (Excoffier and

Schneider 2005) that measures the proportion of genetic variation that occurs within and between populations by

comparing multi-locus allele frequencies.

Fig 5: Schematic diagram of research work

2.4.6 Geo-Spatial Tiger Genome Database: All tiger positive scat information is stored in a database with corresponding GPS location, and a site specific Google

map was developed for all these samples with the following information stored on each identified tiger:

Conventional information (Scat, pugmark, camera trap etc),

GPS coordinates

Microsatellite information genotype for each locus

This database, supported by a dedicated server, is searchable; any unknown sample whether it is collected

opportunistically in the field or has been seized can be searched for a match in this database to find out source

location. There are multiple applications and utility of such a database:

Database System Layout

References:

1. Bajimaya, S., S.R. Bhatta and S.R. Jnawali. 2006. The Greater One-Horned Rhinoceros Conservation Action Plan

for Nepal (2006-2011). Department of National Parks and Wildlife Conservation, Ministry of Forests and Soil

Conservation, The Government of Nepal, Kathmandu. <http://assets.panda.

org/downloads/rhino_action_plan_25aug_06_ low_2.pdf>

2. Bajimaya, S., J.B. Karki, G. Pant, M.N. Kafle, S.R.Jnawali, S. Bajracharya, S. Khaling, S.M. Nepal, S. Lohani,

and U.R. Bhuju. 2007 Tiger Conservation Action Plan for Nepal 2008-2012. Department of National Parks and

Wildlife Conservation, Ministry of Forests and Soil conservation, Government of

Nepal.<http://assets.panda.org/downloads/tiger_ conservation_action_plan.pdf>

3. Barlow, A.C.D., C. McDougal, J.L.D. Smith, B. Gurung, S.R. Bhatta, S. Kumal, B. Mahato, and D.B. Tamang.

2009. Temporal variation in tiger (Panthera tigris) populations and its implications for monitoring. Journal of

Mammalogy 90: 472-478

4. DNPWC. 2004. The Snow Leopard Conservation Action Plan for Nepal. Department of National Parks and

Wildlife Conservation, Ministry of Forests and Soil Conservation, the Government of

Nepal.Kathmandu<http://www.dnpwc.gov.np/The%20Snow%20Leopard%20Conservation%20Action%20

Plan%20for%20Nepal.pdf>

5. Gratwicke, B., J. Seidensticker, M. Shrestha, K. Vermilyue and M. Birnbaum. 2007. Evaluating the performance

of a decade of Save The Tiger Fund's investments to save the world's last wild tigers. Environmental Conservation:

1-11

6. Gurung, B.B.. 2002. Mapping the metapopulation structure of tigers throughout Nepalby establishing a tiger

monitoring network of "village rangers". MSc Thesis. University of Minnesota, Twin Cities

7. Karanth, K.U., J.D. Nichols, J. Seidensticker, E.Dinerstein, J.L. David Smith, C. McDougal, A.J.T. Johnsingh, R. S.

Chundawat and W. Thapar. 2003. Science deficiency in conservation practice: the monitoring of tiger populations in India.

Animal Conservation 6: 141-146

8. Sanderson, E., J. Forrest, C. Loucks, J. Ginsberg, E. Dinerstein, J. Seidensticker, P. Leimgruber, M. Songer, A.

Heydlauff, T.O‘Brien, G. Bryja, S. Klenzendorf and E. Wikramanayake. 2006. Setting Priorities for the

Conservation and Recovery of Wild Tigers: 2005-2015. The Technical Assessment. Wildlife Conservation Society,

World Wildlife Fund, and National Fish and Wildlife Service – Save the Tiger Fund, New York – Washington,

D.C.

9. Schipper, J. et al., 2008. The status of the world‘s land and marine mammals: diversity, threat, and knowledge.

Science 141, 67–77.

10. Harihar, A., B. Pandav, and S.P. Goyal. 2009. Subsampling photographic capture-recapture data of tigers

(Panthera tigris) to minimize closure violation and improve estimate precision: a case study. Population Ecology

51: 471-479

11. Jackson. R.M., J.D. Roe, R. Wangchuk, D.O. Hunter. 2006. Estimating snow leopard population abundance using

photography and capture-recapture techniques. Wildlife Society Bulletin 34: 772-781

12. Karanth, U.K., J.D. Nichols, N.S. Kumar and J.E. Hines. 2006. Assessing tiger population dynamics using

photographic capture-recapture sampling. Ecology 87: 2925-2937

13. Bhagavatula, J. and L. Singh. 2006. Genotyping faecal samples of Bengal tiger Panthera tigris tigris for

population estimation: A pilot study. BMC Genetics 4: 48

14. Cracraft, J., J. Feinstein, J. Vaughn, and K. Helm-Bychowski. 1998. Sorting out tigers (Panthera tigris):

mitochondrial sequences, nuclear inserts, systematics and conservation genetics. Animal Conservation 1: 139-150

15. Fernando, P., G. Polet, N. Foead, L.S. Ng, J.Pastorini, and D.J. Melnick. 2006. Genetic diversity, phylogeny

and conservation of the Javan rhinoceros (Rhinoceros sondaicus). Conservation Genetics 7: 439-448

16. Forbes S.H. and D.K. Boyd. 1997. Genetic structure and migration in native and reintroduced Rocky Mountain

wolf populations. Conservation Biology 11: 1226–1234

17. Holdregger, R. and H.H. Wagner. 2008. Landscape Genetics. Bioscience 58: 199-207

18. Janecka, J.E., R. Jackson, Z. Yuquang, L. Diqiang, B. Munkhtsog, V. Buckley-Beason and W.J. Murphy. 2008a.

Population monitoring of snow leopards using noninvasive collection of scat samples: a pilot study. Animal

Conservation 11: 401-411

19. Miller, C.G., P. Joyce, and L.P. Waits. 2005. A new method for estimating the size of small populations from

genetic mark-recapture data. Molecular Ecology 14: 1991-2005

20. Mills, L.S., J.J. Citta, K.P. Lair, M.K. Schwartz, and D.A. Tallmon. 2000. Estimating animal abundance using

noninvasive DNA sampling: promise and pitfalls. Ecological Applications 10: 283-294

21. Mondol, S., K. U. Karanth, N.S. Kumar, A.M. Gopalaswamy, A. Andheria, and U. Ramakrishnan. 2009.

Evaluation of non-invasive genetic sampling methods for estimating tiger population size. Biological Conservation

142: 2350-2360

22. Onorato, D.P., E.C. Hellgren, R.A. Van Den Bussche, D.L. Doan-Crider and J. Raymon Skiles Jr.2009. Genetic

structure of American black bears in the desert southwest of North America: conservation implications for

recolonization. Conservation Genetics 8: 565-576

23. Singh, A., A. Gaur, K. Shailaja, B.S. Bala, and L. Singh. 2004. A novel microsatellite (STR) marker for forensic

identification of big cats in India.Forensic Science International 141: 143-14

24. Taberlet P. and G. Luikart. 1999. Non-invasive genetic sampling and individual identification.Biological Journal

of the Linnean Society 68: 41-55.

25. Waits, L.P. and D. Paetkau. 2005. Noninvasive genetic sampling tools for wildlife biologists: a review of applications and

recommendations for accurate data collection. Journal of Wildlife Management 69: 1419-1433

26. Wan, Q., and S. Fang. 2003. Application of species-specific polymerase chain reaction in the forensic

identification of tiger species. ForensicScience International 131: 75-78

27. Wetton, J.H., C.S.F. Tsang, C.A. Roney, A.C. Spriggs.2004. An extremely sensitive species-specificARMs PCR

test for the presence of tiger bone DNA. Forensic Science International 140: 139-145

28. Xu, Y.C., B. Li, W.S. Li, S.Y. Bai, Y. Jin, X.P. Li, M.B. Gu, S.Y. Jing, W. Zhang. 2005. Individualization of tiger

by using microsatellites. Forensic Science International 151: 45-51

29. Caniglia, R., E. Fabbri, C. Greco, M. Galaverni, and E. Randi. 2009. Forensic DNA against wildlife poaching:

Identification of a serial wolf killing inItaly. Forensic Science International: Genetics

30. Coghlan, A. 2003. Forensic test fingers rhino poachers. New Scientist 179: 9-9

31. Lorenzini, R. 2005. DNA forensics and the poaching of wildlife in Italy: A case study. Forensic Science

International 153: 218-221

32. Ogden, R., N. Dawnay, R. McEwing. 2009. Wildlife DNA forensics - bridging the gap between conservation

genetics and law enforcement. Forensic Methods in Conservation Research, Endangered Species Research: pp5 <

http://www.intres.com/abstracts/esr/forensic/pp5/>

33. Palsboll, P.K., M. Berube, H.J.Skaug, and C. Raymakers. 2006. DNA registers of legally obtained wildlife and

derived products as means to identify illegal takes. Conservation Biology 20: 1284-1293

34. Singh, A., A. Gaur, K. Shailaja, B.S. Bala, and L. Singh. 2004. A novel microsatellite (STR) marker for forensic

identification of big cats in India.Forensic Science International 141: 143-147

35. Wan, Q., and S. Fang. 2003. Application of species-specific polymerase chain reaction in the forensic identification of tiger

species. ForensicScience International 131: 75-78

36. Qin J, Li R, Raes J, Arumugam M, Burgdorf KS, Manichanh C, Nielsen T, Pons N, Levenez F, Yamada T, et al.:

A human gut microbial gene catalogue established by metagenomic sequencing.Nature 2010, 464(7285):59-65.

37. Gill SR, Pop M, DeBoy RT, Eckburg PB, Turnbaugh PJ, Samuel BS, Gordon JI, Relman DA, Fraser-Liggett CM,

Nelson KE: Metagenomic analysis of the human distal gut microbiome. Science 2006, 312(5778):1355-1359.

38. Hugenholtz P, Tyson G: Microbiology:metagenomics.Nature 2008, 455(7212):481-483

39. Chakravorty S, Helb D, Burday M, Connell N, Alland D: A detailed analysis of 16S ribosomalnRNA gene

segments for the diagnosis of pathogenic bacteria.Journal of microbiologicalmethods 2007, 69(2):330-339

40. HaoX,JiangR,ChenT:Clustering 16S rRNA for OUTprediction: a method of unsupervised Bayesianclustering. Bioinformatics

2011, 27(5):611-618

41. Sun Y, Cai Y, Liu L, Yu F, Farrell M, McKendree W, Farmerie W: ESPRIT: estimatingspecies richness using

largecollectionsof 16S rRNA pyrosequences.NucleicAcidsResearch 2009,37(10):e76-e76

42. Sogin ML, Morrison HG, Huber JA, Welch D, Huse SM, Neal PR, Arrieta JM, Herndl GJ: Microbial diversity in the deep

sea and the underexplored "rare biosphere".Proc Natl AcadSci U S A 2006,103(32):12115-12120

43. Weisburg WG, Barns SM, Pelletier DA, Lane DJ (January 1991). "16S ribosomal DNA amplification for

phylogenetic study". J Bacteriol. 173 (2): 697– 703. PMC 207061.PMID 1987160.

CHAPTER III

Incidental discovery of non-focal carnivore species during genetic study of Bengal tiger (Panthera

tigris tigris) and snow leopard (Panthera uncia) in Nepal

Dibesh Karmacharya1§

, Sulochana Manandhar1, Santosh Dulal

1, Jivan Shakya

1, Kanchan Thapa

2, Manisha Bista

1,

Govind P. Sah1, Ajay N. Sharma

1, Adarsh Sherchan

1, Prajwol Manandhar

1, Laura D. Bertola

3 , Jan Janecka

4,

Marcella Kelly2, Lisette Waits

5

1Center for Molecular Dynamics-Nepal, Thapathali-11, Kathmandu, Nepal

2Department of Fish and Wildlife Conservation, Virginia Tech, Blacksburg, Virginia, USA

3Institute of Environmental Sciences (CML), Leiden University, Leiden, The Netherlands

4Department of Biological Sciences, Duquesne University, Pittsburgh, Pennsylvania, USA

5Laboratory for Ecological, Evolutionary and Conservation Genetics, University of Idaho, USA

§Corresponding author ([email protected])

[This article has been published in the Asian Journal of Conservation Biology; December 2016, Vol.5 no.2; pp 81-88; ISSN

2278-7666]

Abstract

Background

The Bengal tiger (Panthera tigris tigris) and snow leopard (Panthera uncia) are highly endangered apex predators.

These charismatic animals are categorized as flagship and umbrella species, and hence are the focus of many

conservation programs. Protecting tiger and snow leopard also safeguards entire habitat in which they reside,

including other non-focal sympatric carnivores.

Results

In our non-invasive genetic study of the tiger in the Chitwan National Park (CNP) (2011-2013), out of total collected

scat samples (n=420), only 56% (n=237) were of tiger. The remaining non-tiger samples (n=183) included non-focal

carnivores; leopard (Panthera pardus, n=83), leopard cat (Prionailurus bengalensis, n=10), jungle cat (Felis chaus,

n=2), fishing cat (Prionailurus viverrinus, n=2) and fox (Vulpes spp., n=10). Out of total 32 putative snow leopard

scat samples collected from the Mustang region of Annapurna Conservation Area (ACA) (2010-2011), only 56% (n=

18) were found to be snow leopard. We identified red fox (Vulpes vulpes, n=6), Himalayan wolf (Canis spp., n=1),

leopard cat (P. bengalensis, n=1), lynx (Lynx lynx, n=1) and leopard (P. pardus, n=1) from the remaining samples

using DNA barcoding. Remarkably, we were able to document various threatened carnivores categorized as either

critically endangered (wolf), endangered (fishing cat) or vulnerable (lynx, leopard) on Nepal‘s red list mammal status.

The presence of lynx in alpine forest at elevation of 3939 m, Himalayan wolf in rugged Trans-Himalayan terrain at

elevation of 4317 m in the snow leopard habitat were some significant findings from our non-invasive genetic study.

Similarly, the spatial distribution of three small felids: leopard cat, jungle cat and fishing cat in sub-tropical deciduous

forest of Terai has provided insights on sympatric occurrences of small cats among themselves as well as with two

other large felids leopard and tiger in CNP.

Conclusions

Our non-invasive sampling based genetic studies on tiger and snow leopard have detected other non-focal carnivore

species that co-share the same habitat. This hence has significantly increased our understanding of the carnivore

communities in two different landscapes, Terai and Himalayas, of Nepal that are prime habitat for snow leopards and

tigers.

Key Words: Non-invasive, DNA barcoding, sympatric, non-focal carnivore, flagship

Background

Amid concerns over the loss of biodiversity, experts have been increasingly seeking clues to understand ecosystems

and biodiversity [1]. The prevailing conservation and management practices, globally, include coarse-filter and fine-

filter approaches [2], the former aims to preserve entire communities of plants and animals by protecting large extents

of habitat (ecosystem approach) whilst the latter focuses to protect species (species approach).

Nepal established several protected areas based on species and ecosystem approaches for wildlife and habitat

management following prevailing conservation trends in the 1970s [3]. The ecosystem conservation and management

approach has largely been boosted by research on mega species such as the greater one-horned rhinoceros

(Rhinoceros unicornis), Bengal tiger (P. t. tigris, hereafter referred as tiger) and snow leopard (P. uncia).

Maintaining viable populations of these mega species, that include apex predators, is important for ecosystem

integrity and resilience [4]. This has been demonstrated well by successful programs with landscape level approach

to conservation like the Terai Arc Landscape (TAL) project in Nepal, which has recently contributed to the gradual

recovery in the tiger population [5] including other mega fauna such as rhinoceros and elephants [6, 7].

Current burgeoning conservation needs and limited availability of resources often forces conservation managers and

planners to rely on information on the occurrence of apex species to serve as ecosystem health indicators [8]. Tiger in

the TAL and snow leopard in the high Himalaya including the trans-Himalayan region of Nepal are such apex

predators [9].

Tiger and snow leopard are both endangered species [10]. They are also charismatic and umbrella species [11].

Therefore, they are herded as both means for and targets of conservation. By protecting tiger and snow leopard, it is

hoped that the entire community and ecosystem they reside in can be protected. Being charismatic, they also have

traits and qualities that appeal to target audiences to raise funds and awareness for reducing the biodiversity loss [12].

Core tiger populations are fragmented and found mainly in lowland areas across the four protected areas of Parsa,

Chitwan, Bardia and Shuklaphanta and occupy 36% of 26,000 sq. km of the forested landscape of TAL, Nepal [13-

15]. This habitat is also shared with leopard (P. pardus), dhole (Cuon alpinus), sloth bear (Melursus ursinus) and

striped hyena (Hyaena hyaena) within the TAL [16-19]. The trans-Himalayan region of Nepal is prime habitat for

snow leopard covering an area of 20,000 sq. km [20, 9] and including additional carnivore species, such as wolf (C.

lupus) and lynx (L. lynx) [21, 22].

In the past, there have been numerous studies conducted on the status of the flagship species [23, 24, 5]. However,

there has been limited information available on the degree to which the non-focal species have benefited while

working with these flagship species. Use of non-invasive sampling for genetic studies utilizing molecular scatology

techniques [25, 26] primarily on flagship species could be beneficial in identifying the occurrence of the non-focal

species along with the species of interest , thereby supporting to map and document the spatial distribution of species.

Scat (fecal) surveys are often used to detect and monitor elusive and low-density carnivore species [27-29].

However, there is a high degree of misidentification of scats during field sampling, with many scats not belonging to

the target species of interest [29, 28, 30]. Genetic analysis of DNA extracted from scat samples can address this

problem. One common approach is to identify a species of interest using species-specific PCR primers for

identification [31-34] for the selected pools of samples. However, this approach does not provide identification of any

of the non-focused species. For these samples, DNA sequence data can help identify the species: this is the basis of

DNA barcoding for species identification [35].

During our study on tiger in the Chitwan National Park (CNP) in the lowland areas of TAL under Nepal Tiger

Genome Project [36] in 2011-2013 and on snow leopard in the high elevation of the Mustang region of ACA in Nepal

(Fig. 1), there was significant misidentification of collected scats [28]. Using DNA barcoding technique, we were able

to genetically profile non-focal carnivore species sharing tiger and snow leopard habitats, thereby enhancing our

understanding of the carnivore community across the lowland of Terai and the high elevations in Himalayan

landscapes of Nepal. For broader tiger and snow leopard conservation efforts, it is essential that we have landscape

level of ecological information to understand conservation from ecological perspective. This study aims to provide

the better understanding of major fauna in the habitat coexisting with tiger and snow leopard.

Results

Tiger genetic study

Based on tiger specific PCR, we identified 56% (n=237) of the total collected samples (n=420) as tiger. From the

remaining samples (n=183), we identified leopards (P. pardus, n=83) through leopard specific PCR identification.

Out of remaining 100 samples, we were able to identify leopard cat (P. bengalensis, n=10), fishing cat (P. viverrinus,

n=2), jungle cat (F. chaus, n=2) and fox (Vulpes spp., n=10) through sequencing (Table 3). The rest of the samples

(n=76) yielded poor quality sequencing data, possibly due to poor quality of DNA, which could not be used for

species identification. This suggests there was 31% field misidentification of target species (without including non-

identified samples).

Snow leopard genetic study

Based on snow leopard specific PCR, we identified 56% (n=18) of the total samples (n=32) as snow leopards. Out of

the 14 remaining samples, we determined red fox (V. vulpes, n=6), leopard cat (P. bengalensis, n= 1), wolf (C. lupus,

n=1), leopard (P. pardus, n=1) and lynx (L. lynx, n=1) (Table 4). No DNA amplicon of interest (CYT-B) was amplified

from one sample and three samples yielded bad-quality sequencing data (n=4), which could have been most likely due

to the relatively low quality and quantity of DNA. Overall there was ascertained to be 36% field misidentification of

target species (without including non-identified samples).

Discussion

Species distribution studies based solely on scat morphology can be misleading [37] because sympatric carnivore

species often have indistinguishable scat morphologies [38]. The method of species identification through molecular

scatology has proved to be robust for species distribution studies [39, 40]. In most studies, only scats thought to be

from target species based on morphology are collected, but DNA analysis have frequently revealed that many

collected samples originate from non-focal species reported by previous studies [38, 29, 41, 42], as were also found in

our study.

In this study, 31% (tiger) and 36% (snow leopard) field misidentification of target species was higher than 8% and

4% reported by Mondol et al. 2009 [43] and Borthakur et al. 2011 [44] respectively. The reasons behind high

misidentification could be due to difficulty in distinguishing scats of juvenile tigers with adult leopards or juvenile

leopards with adult snow leopards in the field, comparably minimum expertise of field personnel in scat

identification, the opportunistic sample collection strategy and high degree of non-identification of the total collected

samples most likely due to the relatively low quality of DNA in compromised samples like scats. However, our

accuracy of 56% for the target species (snow leopard) was higher than 39% reported in Kanchenjunga Conservation

Area and Shey Phoksundo National Park of Nepal [28] and 33% reported in Gobi desert in Mongolia [45].

While field misidentification can be a disadvantage when investigating specific species, we can make the best out of

all other samples we have collected non-invasively by identifying them through DNA barcoding using mitochondrial

markers for species identification (Table 3 and Table 4) [46]. Genetic study on mtDNA compared to nuclear DNA is

most scientific and reliable method for species identification because there are various genetic markers currently

available within mtDNA region for specific species identification confirmed through previous studies [47]. In the

field sampling of compromised scat samples, the presence of genetic material is in relatively low quality and quantity,

however mtDNA occurs in high copy numbers in each cell and is able to withstand degradation and environmental

challenges, thus increasing chances of DNA amplification for genetic analysis.

This study was able to find occurrences of three felids (leopard, lynx, leopard cat) and two canids (wolf, red fox)

species in trans-himalayan landscape of Upper Mustang, which is prime habitat for snow leopards. Our study has

revealed one of the first strong records of lynx in Nepalese Himalayas using non-invasive genetics, as there is

currently very limited information on the occurrences of this species in Nepal [48]. This study has discerned insights

on sympatry between snow leopards and other large predators such as leopards and wolves in Mustang region of

ACA.

Similarly, our tiger study in CNP has found occurrences of four other felid species, one large (leopard) and three

small (leopard cat, jungle cat, fishing cat) that co-share habitat with tigers. Among canids, we detected a species of

fox demonstrated to have match with red fox (V. vulpes) sequence in GenBank database, but red foxes are not known

to inhabit Terai landscape. We believe that we might have discovered bengal fox (V. bengalensis) that are known to

occur commonly in Terai region of Nepal and India [49], whose reference sequences are not available in GenBank

database. It has also been reported by previous study [29] that this ~100 bp fragment of CYT-B gene is not able to

differentiate between different species of Vulpes, hence additional studies might be required to verify our assumption.

The wildlife genetic study is vital to determine and construct baseline genetic information on species including those

that are currently not focus for studies. By analyzing misidentified samples, we have gathered information on those

species that are under various threatened categories according to national red list assessment [48]. Wolf is categorized

as "Critically Endangered", fishing cat as "Endangered", while leopard cat, lynx and leopard are categorized

"Vulnerable" in the national red list assessment [48]. Among these threatened taxa, wolf, leopard cat and lynx are

also listed as protected mammal species under the National Parks and Wildlife Conservation Act 1973 of Government

of Nepal [50]. Our study has documented lynx in alpine forest habitat at an elevation of 3939 m above sea level of

Mustang region in ACA. In ACA, lynx are thought to have potential distribution across Trans-Himalaya of Upper

Mustang [48], but we have identified from alpine forest near Jomsom. Due to their elusiveness, there is lack of

complete information about their distribution in the Himalayas of Nepal, this record of lynx in alpine forest habitat

denotes that more studies are required to understand the ecology of such charismatic and protected species. Leopard

cats are found in a broad range of habitat types from tropical lowland rain forests to coniferous forests in the

Himalayas [51], their occurrence in our survey sites agrees with this distribution range from sub-tropical deciduous

forest (elevation 150- 800 m) in plains of Terai across CNP to alpine forest (elevation 3876 m) in the high Himalaya

of Mustang. Our claim on the habitat of leopard cat was also supported by the previous study which recorded the

presence of leopard cat at highest elevation (4474 m) across snow leopard habitat at Kanchenjunga Conservation Area

[52]. Wolves are among the least studied species in Nepal, recent genetic studies have provided insights about

divergent taxonomic placement of wolves roaming the Himalayas [53-55]. We think we might have also discovered

the Himalayan wolf in Upper Mustang, but we are not able to confirm it because of shorter sequence fragment that

cannot differentiate between different species of wolves. The wolf scat was detected alongside the scat identified as

snow leopard in rugged Trans-Himalayan terrain at elevation of 4317 m, which is very interesting information from

the point of view of behavioral ecology of these two predators, as carnivore scat deposits are mostly associated with

their territorial marking behavior [56-59].

Among large felids, leopards are considered to be habitat generalist. They are known to occur across Himalayan

regions through Mid-hills to Terai regions of Nepal [48]. Here in our study, we have detected their occurrence at one

instance in Mustang and at multiple instances in CNP sharing habitat with snow leopards and tigers in high Himalayas

and lowland Terai respectively. CNP is one of the few national parks in the world boasting six felid species [60],

among which three are small felids. We have detected all of the three small felids in CNP through our non-invasive

genetic study. Leopard cats are the most frequently captured small felids in our study along with jungle cat and fishing

cat. The spatial data of these three small felids across CNP provides vital understanding about their sympatric

occurrences in the same habitat. Fishing cat is a rare species known to have sparse distribution range in Terai plains of

Nepal and is known to highly associate with wetland areas [61-63]. Here, we have detected two instances of fishing

cat, both of which fall under their known distribution range in CNP and lie very close to river (< 2 km). Fishing cats

have been recorded outside protected areas in Jagdishpur Reservior, a Ramsar site lying along a rural settlement area

of south-western Nepal [64]. One of the instances of fishing cat in our study was also found to be located in distance

less than 100 m from nearest human settlement of Amaltari village in Chitwan.

Conclusion

In the on-going debate on whether focusing solely on an umbrella and/or flagship species is beneficial [65], this study

revealed that even while conducting research on focal carnivore species such as tiger or snow leopard, there is ample

opportunity to collect data on other, often neglected, non-focal carnivore species that co-occur in the same habitat.

This emphasizes the importance of holistic sampling and monitoring approaches that capture broader ecological

information. This non-invasive genetic sampling approach, integrated in our studies of tiger and snow leopard, has

significantly increased our understanding on sympatric coexisting carnivores and provided us with the vital

information about their rough distribution. Hence, this study has in overall contributed expanding our knowledge of

the carnivore community in the Terai and Himalaya regions of Nepal.

Methods

Putative tiger scat samples (n=420) were collected from CNP in 2011-2013. Samples were opportunistically collected

across selected routes (fire-lines, trails and riverbanks) within core areas of CNP (Fig. 1). Similarly, putative snow

leopard scat samples (n= 48) were opportunistically collected in 2010-2011 from Mustang region of ACA. The study

sites were selected based on suitable habitat and recently reported high activities for snow leopard (Fig. 1).

Collected samples were stored in DET buffer (20% DMSO, 0.25 M Sodium-EDTA, NaCl to saturation, pH 7.5) and

transported in ambient temperature to the laboratory of the Center for Molecular Dynamics Nepal in Kathmandu.

Total DNA was extracted from all scat samples using commercially available QIAamp DNA stool mini kit

(QIAGEN, Inc., Germany) following the manufacturer‘s instructions. Extracted DNA was stored at -20°C and

handled carefully to avoid any cross contamination between samples.

Species Identification: tiger, snow leopard and leopard

Species identification on samples was done by Polymerase Chain Reaction (PCR) using species-specific primers (SSP)

for tiger, snow leopard and leopard. For samples from CNP, samples were first addressed to tiger specific primers

(TIF/TIR) [66], which targets ~162 bp fragment of mitochondrial cytochrome b gene. Similarly, for samples from

ACA, samples were first addressed to snow leopard-specific primers (CYTB-SCT-PUN-F/CYTB-SCT-PUN-R) [67],

which targets ~150 bp fragment of mitochondrial cytochrome b gene. The negative samples from these two different

PCR reactions were then addressed to leopard-specific primers (NADH4-F/NADH4-R) [68] that targets ~130 bp

fragment of mitochondrial ND4 gene.

PCR reactions were carried out in a final volume of 7 µl containing 3.5 µl Qiagen mastermix (2X), 0.7 µl Q-solution,

0.07 µl of each forward and reverse primer at a final concentration of 0.2 µM, 0.66 µl nuclease free water and 2 µl

template DNA for tiger samples and 1.16 µl nuclease free water and 1.5 µl template DNA for snow leopard samples,

respectively. DNA from known tiger and snow leopard tissue samples generously provided by Dr. Jan Janecka, was

used as a positive control. PCR reactions were carried out in the thermocycler (MJ Research Tetrad PTC-225

Thermal Cycler, USA) with the conditions shown in Table 2. The amplified PCR products (~162 bp for tiger, ~150

bp for snow leopard and ~130 bp for leopard) were visualized on 1.8 % agarose gel electrophoresis.

General carnivore identification

The species barcoding of samples that were not identified as tigers, snow leopards and leopards were performed by

amplifying short fragment (~146 bp) of cytochrome b gene and sequencing the PCR product. The PCR was done

using a unique primer set (CYTB-SCT-F/CYTB-SCT-R) [38] which amplifies a short variable region of CYT-B in

carnivores. PCR was performed in a final 25 µl reaction mixture containing 12.5 µl Qiagen mastermix, 2.5 µl Q-

solution, 0.63 µl each forward and reverse primers at a final concentration of 0.5 µM, 6.74 µl nuclease free water and

2 µ DNA. PCR reactions were carried out in MJ Research PTC-225 Thermal Cycler, USA with the conditions as

shown in Table 2. The PCR products, along with incorporated known carnivore sample as positive control, were

visualized on 2% agarose gel. The carnivore positive amplicons were sequenced on an ABI 310 machine using the

forward primer (CYTB-SCT-F). The DNA sequences were edited and compared with reference GenBank database

using BLAST tool in NCBI to identify carnivore species. The species identification was confirmed following the

criteria of ≥ 90% query coverage and ≥ 97% identity.

Acknowledgements

We would like to express our deepest gratitude to Dr. Rodney Jackson of the Snow Leopard Conservancy for

supporting us in many ways, including providing us with funds for the study. We would like to thank the entire team

of NTGP, including Waits Laboratory for providing training and supporting us with this study. USAID/Nepal

provided funding for NTGP, and Mr. Netra Sharma of USAID/Nepal helped in NTGP project implementation. We

would also like to thank the Department of National Parks and Wildlife Conservation and Ministry of Forests and Soil

Conservation (Nepal) for their continuing support in our conservation genetics work in Nepal and providing us with

the necessary permits. Mr. Pema Lama and his field team assisted us in collecting samples from Mustang area. This

study was also supported by the Snow Leopard Network through its Snow Leopard Conservation Grants.

References

1. Cardinale BJ, Duffy JE, Gonzalez A, Hooper DU, Perrings C, Venail P et al. Biodiversity loss and its impact on

humanity. Nature. 2012;486(7401):59-67. doi:10.1038/nature11148.

2. Noss RF. From plant communities to landscapes in conservation inventories: a look at The Nature Conservancy

(USA). Biological Conservation. 1987;41(1):11-37.

3. Heinen JT, Shrestha SK. Evolving policies for conservation: an historical profile of the protected area system of

Nepal. Journal of Environmental Planning and Management. 2006;49(1):41-58.

4. Soulé ME, Terborgh JW. Continental conservation: scientific foundations of regional reserve networks. Island

Press; 1999.

5. GoN. Summary Report : Status of Tiger and Prey-Base Population in Nepal. Kathmandu, Nepal., Department of

National Parks and Soil Conservation DoF;2013 2013.

6. Thapa K, Nepal S, Thapa G, Bhatta SR, Wikramanayake E. Past, present and future conservation of the greater

one-horned rhinoceros Rhinoceros unicornis in Nepal. Oryx. 2013;47(03):345-51.

7. GoN. Elephant Conservation Action Plan for Nepal., Department of National Parks and Soil Conservation

DoF;2009.

8. Simberloff D. Flagships, umbrellas, and keystones: Is single-species management passé in the landscape era?

Biological conservation. 1998;83(3):247-57. doi:http://dx.doi.org/10.1016/S0006-3207(97)00081-5.

9. Forrest JL, Wikramanayake E, Shrestha R, Areendran G, Gyeltshen K, Maheshwari A et al. Conservation and

climate change: Assessing the vulnerability of snow leopard habitat to treeline shift in the Himalaya. Biological

Conservation. 2012;150(1):129-35.

10. The IUCN Red List of Threatened Species. Version 2013.1 http://www.iucnredlist.org. Downloaded on 24

September 2013. [database on the Internet]2013. Accessed: 24 September 2013

11. Clucas B, McHugh K, Caro T. Flagship species on covers of US conservation and nature magazines. Biodiversity

and Conservation. 2008;17(6):1517-28.

12. Verissimo D, Pongiluppi T, Santos MC, Develey PF, Fraser I, Smith RJ et al. Using a Systematic Approach to

Select Flagship Species for Bird Conservation. Conservation biology : the journal of the Society for Conservation

Biology. 2013. doi:10.1111/cobi.12142.

13. Shrestha N. Protected wildlife species of Nepal: an introductory handbook. IUCN; 1997.

14. Smith JLD, Ahern SC, McDougal C. Landscape analysis of tiger distribution and habitat quality in Nepal.

Conservation Biology. 1998;12(6):1338-46.

15. Barber‐Meyer S, Jnawali S, Karki J, Khanal P, Lohani S, Long B et al. Influence of prey depletion and human

disturbance on tiger occupancy in Nepal. Journal of Zoology. 2013;289(1):10-8.

16. Odden M, Wegge P, Fredriksen T. Do tigers displace leopards? If so, why? Ecological research. 2010;25(4):875-

81.

17. Joshi AR, Garshelis DL, Smith JL. Home ranges of sloth bears in Nepal: implications for conservation. The

Journal of wildlife management. 1995:204-14.

18. Seidensticker J. On the ecological separation between tigers and leopards. Biotropica. 1976:225-34.

19. Thapa K, Kelly MJ, Karki JB, Sub‐edi N. Distribution update. 2013.

20. Jackson R, Ahlborn G. Snow leopards (Panthera uncia) in Nepal: home range and movements. National

Geographic Research. 1989;5(2):161-75.

21. Mishra C, Allen P, McCarthy T, Madhusudan M, Bayarjargal A, Prins HH. The role of incentive programs in

conserving the snow leopard. Conservation Biology. 2003;17(6):1512-20.

22. Namgail T, Fox JL, Bhatnagar YV. Carnivore-caused livestock mortality in Trans-Himalaya. Environmental

management. 2007;39(4):490-6. doi:10.1007/s00267-005-0178-2.

23. Subedi N, Jnawali SR, Dhakal M, Pradhan N, Lamichhane BR, Malla S et al. Population status, structure and

distribution of the greater one-horned rhinoceros Rhinoceros unicornis in Nepal. Oryx. 2013;47(03):352-60.

24. Karki JB, Pandav B, Jnawali SR, Shrestha R, Pradhan NMB, Lamichane BR et al. Estimating the abundance of

Nepal's largest population of tigers Panthera tigris. Oryx. 2013;FirstView:1-7. doi:doi:10.1017/S0030605313000471.

25. Kelly MJ, Betsch J, Wultsch C, Mesa B, Mills LS. Noninvasive sampling for carnivores. Carnivore ecology and

conservation: a hand-book of techniques Oxford University Press, Oxford. 2012:47-67.

26. Reed J, Tollit D, Thompson P, Amos W. Molecular scatology: the use of molecular genetic analysis to assign

species, sex and individual identity to seal faeces. Molecular ecology. 1997;6(3):225-34.

27. Waits LP, Paetkau D. Noninvasive genetic sampling tools for wildlife biologists: a review of applications and

recommendations for accurate data collection. Journal of Wildlife Management. 2005;69(4):1419-33.

28. Karmacharya DB, Thapa K, Shrestha R, Dhakal M, Janecka JE. Noninvasive genetic population survey of snow

leopards (Panthera uncia) in Kangchenjunga conservation area, Shey Phoksundo National Park and surrounding

buffer zones of Nepal. BMC research notes. 2011;4:516. doi:10.1186/1756-0500-4-516.

29. Janečka J, Jackson R, Yuquang Z, Diqiang L, Munkhtsog B, Buckley‐Beason V et al. Population monitoring of

snow leopards using noninvasive collection of scat samples: a pilot study. Animal Conservation. 2008;11(5):401-11.

30. Spiering PA, Gunther MS, Wildt DE, Somers MJ, Maldonado JE. Sampling error in non-invasive genetic

analyses of an endangered social carnivore. Conservation Genetics. 2009;10(6):2005-7.

31. Fernandes CA, Ginja C, Pereira I, Tenreiro R, Bruford MW, Santos-Reis M. Species-specific mitochondrial DNA

markers for identification of non-invasive samples from sympatric carnivores in the Iberian Peninsula. Conservation

genetics. 2008;9(3):681-90.

32. Kurose N, Masuda R, Tatara M. Fecal DNA analysis for identifying species and sex of sympatric carnivores: a

noninvasive method for conservation on the Tsushima Islands, Japan. The Journal of heredity. 2005;96(6):688-97.

doi:10.1093/jhered/esi124.

33. Nagata J, Aramilev VV, Belozor A, Sugimoto T, McCullough DR. Fecal genetic analysis using PCR-RFLP of

cytochrome b to identify sympatric carnivores, the tiger Panthera tigris and the leopard Panthera pardus, in far eastern

Russia. Conservation Genetics. 2005;6(5):863-6.

34. Oliveira R, Castro D, Godinho R, Luikart G, Alves P. Species identification using a small nuclear gene fragment:

application to sympatric wild carnivores from South-western Europe. Conservation Genetics. 2010;11(3):1023-32.

35. Baselga A, Fujisawa T, Crampton-Platt A, Bergsten J, Foster PG, Monaghan MT et al. Whole-community DNA

barcoding reveals a spatio-temporal continuum of biodiversity at species and genetic levels. Nature communications.

2013;4:1892. doi:10.1038/ncomms2881.

36. NTGP/CMDN. Nepal Tiger Genome Project - Final Report 2014, Center for Molecular Dynamics-Nepal: Center

for Molecular Dynamics-Nepal2014.

37. Davison A, Birks JDS, Brookes RC, Braithwaite TC, Messenger JE. On the origin of faeces: morphological

versus molecular methods for surveying rare carnivores from their scats. Journal of Zoology. 2002;257(2):141-3.

doi:10.1017/s0952836902000730.

38. Farrell LE, Roman J, Sunquist ME. Dietary separation of sympatric carnivores identified by molecular analysis of

scats. Mol Ecol. 2000;9(10):1583-90. doi:10.1046/j.1365-294x.2000.01037.x.

39. Palomares F, Godoy JA, Piriz A, O'Brien SJ. Faecal genetic analysis to determine the presence and distribution of

elusive carnivores: design and feasibility for the Iberian lynx. Mol Ecol. 2002;11(10):2171-82.

40. Cossíos ED, Madrid A, Condori JL, Fajardo U. Update on the distribution of the Andean cat Oreailurus jacobita

and the pampas cat Lynchailurus colocolo in Peru. 2007.

41. Sugimoto T, Nagata J, Aramilev VV, McCullough DR. Population size estimation of Amur tigers in Russian Far

East using noninvasive genetic samples. Journal of Mammalogy. 2012;93(1):93-101. doi:10.1644/10-mamm-a-368.1.

42. Perez I, Geffen E, Mokady O. Critically Endangered Arabian leopards Panthera pardus nimr in Israel: estimating

population parameters using molecular scatology. Oryx. 2006;40(03):295-301. doi:doi:10.1017/S0030605306000846.

43. Mondol S, Ullas Karanth K, Samba Kumar N, Gopalaswamy AM, Andheria A, Ramakrishnan U. Evaluation of

non-invasive genetic sampling methods for estimating tiger population size. Biological Conservation.

2009;142(10):2350-60.

44. Borthakur U, Barman RD, Das C, Basumatary A, Talukdar A, Ahmed MF et al. Noninvasive genetic monitoring

of tiger (Panthera tigris tigris) population of Orang National Park in the Brahmaputra floodplain, Assam, India.

European Journal of Wildlife Research. 2011;57(3):603-13.

45. Janecka JE, Munkhtsog B, Jackson RM, Naranbaatar G, Mallon DP, Murphy WJ. Comparison of noninvasive

genetic and camera-trapping techniques for surveying snow leopards. Journal of Mammalogy. 2011;92(4):771-83.

46. Madden T. The BLAST sequence analysis tool. The NCBI Handbook [Internet] National Library of Medicine

(US), National Center for Biotechnology Information, Bethesda, MD. 2002.

47. Alacs EA, Georges A, FitzSimmons NN, Robertson J. DNA detective: a review of molecular approaches to

wildlife forensics. Forensic science, medicine, and pathology. 2010;6(3):180-94. doi:10.1007/s12024-009-9131-7.

48. Jnawali S, Baral H, Lee S, Acharya K, Upadhyay G, Pandey M et al. The Status of Nepal Mammals: The National

Red List Series, Department of National Parks and Wildlife Conservation Kathmandu, Nepal. Preface by Simon M

Stuart Chair IUCN Species Survival Commission The Status of Nepal‘s Mammals: The National Red List Series.

2011:4.

49. Gompper ME, Vanak AT. Vulpes bengalensis. Mammalian Species. 2006:1-5.

50. NPWCA. National Parks and Wildlife Conservation Act. 1973. National Parks and Wildlife Conservation Act

2029. Government of Nepal.1973.

51. Mohamed A, Sollmann R, Bernard H, Ambu LN, Lagan P, Mannan S et al. Density and habitat use of the leopard

cat (Prionailurus bengalensis) in three commercial forest reserves in Sabah, Malaysian Borneo. Journal of

Mammalogy. 2013;94(1):82-9.

52. Thapa K, Pradhan N, Barker J, Dhakal M, Bhandari A, Gurung G et al. High elevation record of a leopard cat in

the Kangchenjunga Conservation Area, Nepal. Cat News. 2013;58:26-7.

53. Chetri M, Jhala YV, Jnawali SR, Subedi N, Dhakal M, Yumnam B. Ancient Himalayan wolf (Canis lupus

chanco) lineage in Upper Mustang of the Annapurna Conservation Area, Nepal. ZooKeys. 2016(582):143.

54. Aggarwal RK, Kivisild T, Ramadevi J, Singh L. Mitochondrial DNA coding region sequences support the

phylogenetic distinction of two Indian wolf species. Journal of Zoological Systematics and Evolutionary Research.

2007;45(2):163-72.

55. Sharma DK, Maldonado JE, Jhala YV, Fleischer RC. Ancient wolf lineages in India. Proceedings of the Royal

Society of London B: Biological Sciences. 2004;271(Suppl 3):S1-S4.

56. Barja I, de Miguel FJ, Bárcena F. The importance of crossroads in faecal marking behaviour of the wolves (Canis

lupus). Naturwissenschaften. 2004;91(10):489-92.

57. Barja I, de Miguel FJ, Bárcena F. Faecal marking behaviour of Iberian wolf in different zones of their territory.

Folia Zoologica. 2005;54(1/2):21.

58. Piñeiro A, Barja I. The plant physical features selected by wildcats as signal posts: an economic approach to fecal

marking. Naturwissenschaften. 2012;99(10):801-9.

59. Piñeiro A, Barja I. Evaluating the function of wildcat faecal marks in relation to the defence of favourable hunting

areas. Ethology Ecology & Evolution. 2015;27(2):161-72.

60. Lamichhane B, Dhakal M, Subedi N, Pokheral C. Clouded leopard co-exist with other five felids in Chitwan

National Park, Nepal. Cat News. 2014;61:30-2.

61. Jha S. Status and conservation of lowland Terai wetlands in Nepal. Our Nature. 2009;6(1):67-77.

62. Taylor IR, Baral HS, Pandey P, Kaspal P. The conservation status of the Fishing Cat Prionailurus viverrinus

Bennett, 1833 (Carnivora: Felidae) In Koshi Tappu Wildlife Reserve, Nepal. Journal of Threatened Taxa.

2016;8(1):8323-32.

63. Mukherjee S, Appel, A., Duckworth, J.W., Sanderson, J., Dahal, S., Willcox, D.H.A., Herranz Muñoz, V., Malla,

G., Ratnayaka, A., Kantimahanti, M., Thudugala, A., Thaung R. & Rahman, H. . Prionailurus viverrinus. The IUCN

Red List of Threatened Species 2016: e.T18150A50662615. Downloaded on 15 February 2017. 2016.

64. Dahal S, Baral S, Nepal M, Neupane KR, Dahal BV, Kathmandu B. Status of Fishing Cat in Jagadishpur

Reservoir and Ghodaghodi Lake and assessment of threat. 2014.

65. Andelman SJ, Fagan WF. Umbrellas and flagships: efficient conservation surrogates or expensive mistakes?

Proceedings of the National Academy of Sciences of the United States of America. 2000;97(11):5954-9.

doi:10.1073/pnas.100126797.

66. Bhagavatula J, Singh L. Genotyping faecal samples of Bengal tiger Panthera tigris tigris for population

estimation: a pilot study. BMC genetics. 2006;7(1):48.

67. Janecka JE, Jackson R, Munkhtsog B, Murphy WJ. Characterization of 9 microsatellites and primers in snow

leopards and a species-specific PCR assay for identifying noninvasive samples. Conservation Genetics Resources.

2014;6(2):369-73.

68. Mondol S, Navya R, Athreya V, Sunagar K, Selvaraj V, Ramakrishnan U. A panel of microsatellites to

individually identify leopards and its application to leopard monitoring in human dominated landscapes. BMC

genetics. 2009;10(1):79.

Table 1. PCR primers used in this study

Species

identification Primers Primer sequence (5'-3') References

Tiger TIF

TIR

5'-ATAAAAAATCAGGAATGGTG-3'

5'-TGGCGGGGATGTAGTTATCA-3' [66]

Snow leopard CYTB-SCT-PUN-F

CYTB-SCT-PUN-R

5'-TGGCTGAATTATCCGATACC-3'

5'-AGCCATGACTGCGAGCAATA-3' [67]

Leopard NADH4-F

NADH4-R

5'-TRATAGCTGCYTGATGAC-3'

5'-GTTTGTGCCTATAAGGAC-3' [68]

Carnivore CYTB-SCT- F

CYTB-SCT- R

5'-AAACTGCAGCCCCTCAGAATGATATTTGTCCTCA-3'

5'-TATTCTTTATCTGCCTATACATRCACG-3' [38]

Table 2. Species specific PCR conditions used for the amplication of gene of interest in this study

Species specific

PCR

Temperature (°C) Time Cycles Gene of interest

(Amplicon size) Tiger 95 15 min 1 CYT-B, ~162 bp

94 30 sec 35

59 90 sec

72 90 sec

72 10 min 1

Snow leopard 95 15 min 1 CYT-B, ~150 bp

95 30 sec 45

60 15 sec

72 60 sec

72 10 min 1

Leopard 95 15 min 1 NADH4, ~130 bp

94 15 sec 50

56 30 sec

72 30 sec

72 10 min 1

Carnivore 95 15 min 1 CYT-B, ~146 bp

94 30 sec 45

55 30 sec

72 60 sec

Table 3. NCBI BLAST results of target amplification of Cytochrome b (CYT-B) region of mtDNA in non-tiger

samples.

Sample

code

Query

length

(bp)

GenBank

ACC

Species Common

name

Query

Coverage

(%)

Identity

(%)

NTGP-590 90 KP202274 Felis chaus jungle cat 100 97.78

NTGP-643 73 EU362125 Felis chaus jungle cat 100 98.63

NTGP-872 88 KP246843 Prionailurus bengalensis leopard cat 100 100

NTGP-539 88 KP246843 Prionailurus bengalensis leopard cat 100 100

NTGP-875 80 KP246843 Prionailurus bengalensis leopard cat 100 98.75

NTGP-879 91 KP246843 Prionailurus bengalensis leopard cat 100 98.9

NTGP-606 91 KP246843 Prionailurus bengalensis leopard cat 100 96.7

NTGP-886 87 KP246843 Prionailurus bengalensis leopard cat 100 98.85

NTGP-870 88 KP246843 Prionailurus bengalensis leopard cat 100 98.86

NTGP-586 92 KP246843 Prionailurus bengalensis leopard cat 100 98.91

NTGP-601 88 KP246843 Prionailurus bengalensis leopard cat 100 100

NTGP-871 88 KP246843 Prionailurus bengalensis leopard cat 100 98.86

NTGP-584 80 KP202270 Prionailurus viverrinus fishing cat 100 100

NTGP-616 90 KP202270 Prionailurus viverrinus fishing cat 100 100

NTGP-468 95 KT448287 Vulpes vulpes red fox 100 96.84

NTGP-632 66 KT448287 Vulpes vulpes red fox 100 96.97

NTGP-296 83 KT448287 Vulpes vulpes red fox 97.59 98.77

NTGP-327 88 KT448287 Vulpes vulpes red fox 96.59 100

NTGP-469 62 KT448287 Vulpes vulpes red fox 90.32 100

NTGP-604 82 KT448287 Vulpes vulpes red fox 96.34 100

NTGP-651 92 KT448287 Vulpes vulpes red fox 100 97.83

NTGP-1024 87 KT448287 Vulpes vulpes red fox 96.55 98.81

NTGP-1026 92 KT448287 Vulpes vulpes red fox 94.57 100

NTGP-608 80 KT448287 Vulpes vulpes red fox 96.25 100

Table 4. NCBI BLAST results of target amplification of Cytochrome b region of mtDNA in non-snow leopard

samples.

sample

code

Query

length (bp)

GenBank ACC Species Common

name

Query

coverage

(%)

Identity

(%)

W-SL-96 115 KT448287 Vulpes vulpes red fox 99.13 98.25

W-SL-97 101 KT448287 Vulpes vulpes red fox 100 94.06

W-SL-99 119 KT448287 Vulpes vulpes red fox 100 97.48

W-SL-100 120 KT448287 Vulpes vulpes red fox 99.17 97.48

W-SL-101 115 KT448287 Vulpes vulpes red fox 99.13 98.25

W-SL-111 120 KT901460 Canis lupus wolf 100 97.5

W-SL-114 111 EU434732 Prionailurus

bengalensis

leopard

cat

100 100

W-SL-118 121 EF056507 Panthera pardus leopard 100 98.35

W-SL-119 105 KT448287 Vulpes vulpes red fox 99.05 97.12

W-SL-121 95 KP202283 Lynx lynx lynx 100 98.95

Figure 1. Map of the study sites. Chitwan National Park (CNP) located in the plains of Terai. Tiger tends to occupy

the lowland area Terai Arc Landcape (TAL) of Nepal. Mustang is located within the Annapurna Conservation Area

(ACA). Snow leopards are known to inhabit throughout trans-Himalayan range of Nepal. Locations showing the

distribution of scat samples and identified species from the study sites in CNP and Mustang Region.

Figure 2. Tiger species identification with the amplification of tiger specific CYT-B using PCR. The tiger-

specific PCR primers amplified a ~162 bp amplicon of Cytochrome b region (CYT-B) of mtDNA. M = 100 bp DNA

ladder (Invitrogen, USA); Lanes 1, 2, 3 & 4 = duplicate tiger samples (W-TG-048, W-TG-049 respectively) tested,

Lanes 5, 6, 7 & 8 = duplicate non-tiger samples (W-TG-50, W-TG-51 respectively) tested; Lane 9 = Negative Control

(NC) = no template DNA; Lane 10 = Positive Control [34] = known tiger DNA

Figure 3. Snow leopard species identification with the amplification of snow leopard specific CYT-B using

PCR. The snow leopard specific PCR primers amplified a ~150 bp amplicon of CYT-B of mtDNA. M = 100 bp DNA

ladder (Invitrogen, USA); Lanes 1, 2, 3, 4, 11, 12, 13 & 14 = duplicate snow leopard samples (W-SL-1313, W-SL-

1316, W-1320, W-SL-21 respectively) tested, Lanes 5, 6, 7, 8, 9 & 10 = duplicate non-snow leopard samples (W-SL-

1317, W-SL-1318, W-SL-1319 respectively) tested; Lane 15 (NC) = no template DNA and Lane 16 [34] = known

snow leopard DNA.

Figure 4. Leopard species identification with the amplification of leopard specific NADH4 using PCR. The

leopard specific PCR primers amplified a 130 bp NADH4 amplicon of mtDNA. M = 100 bp DNA ladder (Invitrogen,

USA); Lanes 1, 2, 7 & 8 = duplicate non-snow leopard samples (W-TG-016, W-TG-031 respectively) tested, Lanes 3,

4, 5 & 6 = duplicate snow leopard samples (W-TG-021, W-TG-022) tested; Lane 9 (NC) = no template DNA and

Lane 10 [34] = known snow leopard DNA.

Figure 5. Carnivore identification using PCR and DNA sequencing of carnivore specific CYT-B. Carnivore

specific primers amplified ~146 bp CYT-B amplicon. M = 100 bp DNA ladder (Invitrogen, USA); Lanes 1, 2, 3, 4, & 5

= carnivore samples (W-SL-96, W-SL-114, W-SL-121, NTGP-616, NTGP-643 respectively) tested, Lane 6 (NC) = no

template DNA and Lane 7 [34] = known carnivore DNA

CHAPTER IV

Assessment of genetic diversity, population structure, and gene flow of tigers (Panthera tigris tigris) across

Nepal's Terai Arc Landscape

Short Title:

Genetic diversity assessment of tigers in Nepal

Kanchan Thapa1, #a¶

, Sulochana Manandhar2, Manisha Bista

2, Jivan Shakya

2, Govind Sah

2, Maheshwar Dhakal

3, Netra

Sharma4, Bronwyn Llewellyn

4, Claudia Wultsch

5, Lisette P. Waits

6, Marcella J. Kelly

1, Jean-Marc Hero

7, Jane

Hughes7, Dibesh Karmacharya

2,7¶*

1Department of Fish and Wildlife Conservation, Virginia Tech, Blacksburg, Virginia, USA

#a Current address: Conservation Science Unit, WWF Nepal, Kathmandu, Nepal

2Center for Molecular Dynamics Nepal, Thapathali-11, Kathmandu, Nepal

3Department of National Parks and Wildlife Conservation, Kathmandu, Nepal

4Environment Team, U.S. Agency for International Development, Kathmandu, Nepal

5 American Natural History Museum, New York City, New York, USA

6Department of Fish and Wildlife Sciences, University of Idaho, Moscow, USA

7School of Environment, Griffith University, Queensland, Australia

* Corresponding author

E-mail:[email protected] (DK)

¶ This author contributed equally to this work

[This article has been accepted in the journal PLOS One, and will be published on March 21, 2018]

Abstract

With fewer than 200 tigers (Panthera tigris tigris) left in Nepal, that are generally confined to five protected areas

across the Terai Arc Landscape, genetic studies are needed to provide crucial information on diversity and

connectivity for devising an effective country-wide tiger conservation strategy. As part of the Nepal Tiger Genome

Project, we studied landscape change, genetic variation, population structure, and gene flow of tigers across the Terai

Arc Landscape by conducting Nepal‘s first comprehensive and systematic scat-based, non-invasive genetic survey.

Of the 770 scat samples collected opportunistically from five protected areas and six presumed corridors, 412 were

tiger (57%). Out of ten microsatellite loci, we retain eight markers that were used in identifying 78 individual tigers.

We used this dataset to examine population structure, genetic variation, contemporary gene flow, and potential

population bottlenecks of tigers in Nepal. We detected three genetic clusters consistent with three demographic sub-

populations and found moderate levels of genetic variation (He = 0.61, AR = 3.51) and genetic differentiation (FST =

0.14) across the landscape. We detected 3-7 migrants, confirming the potential for dispersal-mediated gene flow

across the landscape. We found evidence of a bottleneck signature likely caused by large-scale land-use change

documented in the last two centuries in the Terai forest. Securing tiger habitat including functional forest corridors is

essential to enhance gene flow across the landscape and ensure long-term tiger survival. This requires cooperation

among multiple stakeholders and careful conservation planning to prevent detrimental effects of anthropogenic

activities on tigers.

Keywords: Tiger, Genetic variation, Terai arc landscape, Genetic structure, Bottleneck, Gene flow, Noninvasive

genetic sampling

Introduction

Reduction in prime habitat and loss of genetic diversity are influential factors leading to the extirpation of wildlife

populations [1] and extinction of species [2-4]. In the face of habitat fragmentation and isolation, maintaining genetic

connectivity across fragmented landscapes is challenging [5, 6] yet necessary to avert the negative consequences of

genetic drift and inbreeding [7-10]. Maintenance of genetic diversity and gene flow is particularly critical for large

carnivores, which often occur at naturally low densities [11], thus increasing their risk of extinction due to a greater

susceptibility to stochastic events [12, 13].

The tiger (Panthera tigris) is a species of global conservation concern as its range has declined more than 90% in past

two decades, and the species now occupies only 7% of its historic range [14]. The few remaining tigers (est. 3,200)

are concentrated within 76 global Tiger Conservation Landscapes (TCLs) spread across a wide range of tiger habitat

types [14]. Besides poaching, habitat loss and fragmentation remain the largest threats to the survival of extant tiger

populations [15], and threats are mounting in these TCLs [14, 16]. One key region for global tiger conservation is the

Terai Arc Landscape (TAL),which occupies a significant portion (~29,100 km2) of the Eastern Himalayan eco-region

[17] and includes 15 core tiger clusters (identified as Protected Areas) connected by contiguous forest blocks in Nepal

and northwest India [18]. Five protected areas with varying degrees of structural connectivity are located in Nepal,

and are spread in a somewhat linear configuration across primarily forested habitat containing Nepal‘s only tiger

populations (N=198) [19]. Maintaining functional connectivity for tigers in this region is crucial for preserving

genetic variation and long-term population viability [20, 21]. Previous studies of genetic diversity and structure

among tigers have shown that Bengal tigers (Panthera tigris tigris) are the most diverse globally and represent half of

the extant genetic diversity in the species [22]. Multiple studies in India detected moderate to high levels of genetic

diversity and varying levels of gene flow between the tiger sub-populations living in fragmented and human-altered

landscapes [22-27]. Three possible demographic sub-populations of tigers have been identified across the Terai Arc

Landscape of Nepal based on long-term field data and tiger habitat requirements [17, 28], and some degree of

functional connectivity is expected across this region [29]. A recent study has shown that tigers occupy 36% of the

TAL in Nepal and that core tiger populations occur within the protected areas [30]. Few signs of tigers have been

detected outside of protected areas [29]. However, radio-telemetry studies have shown that tigers in the Terai have

dispersed as far as 30 km [31], including through human dominated areas within the landscape [25, 26], and camera

trap data have confirmed the presence of tigers in corridors [32]. The TAL has experienced significant landuse

changes in the recent past [33] that might impede dispersal and gene flow across the landscape and create genetic

subdivision. Thus, a landscape-level genetic study is needed to assess whether dispersal and subsequent breeding

(genetic migration) of tigers occurs across this fragmented region.

We conducted a forest change analysis in the Nepalese portion of the TAL (Fig. 1) to identify major changes in land

use (forest to agriculture) and implemented the first comprehensive fecal based non-invasive genetic assessment of

tigers across the country. We focused our field sampling in 5 protected areas and six putative forest corridors, which

represent the core area for tigers within the TAL-Nepal. Our main study objectives were to: 1) document the scale and

distribution of land use change over the last 300 years in the TAL landscape, 2) determine the number of genetic

groups of tigers, 3) assess genetic diversity, population structure and gene flow, 4) determine the level of

contemporary migration between genetic groups, and 5) test for evidence of population bottlenecks. We hypothesized

that tigers would group into three genetic clusters representing the three demographic tiger populations previously

identified in TAL-Nepal [34]. Given the high degree of land use change in the past, we expected low to moderate

levels of genetic diversity; and limited but detectable gene flow within the region.

Fig 1. Study area and sampling grids cells (15 X 15 km) used for collection of tiger scat samples across the

Terai Arc Landscape, Nepal. We searched and collected tiger scats (n = 770) from 54 grid cells (Protected

Areas: 36 grids; Corridors: 18 grids) out of total 108 grid cells totaling 9,000 km2 of land area.

Materials and Methods

Ethics Statement

All the tiger scat samples were collected non-invasively without capturing or handling any animal. The study permit

was granted by the Department of National Parks and Wildlife Conservation, Ministry of Forest and Soil

Conservation, Government of Nepal, Kathmandu, Nepal (Letter No. 403-08/09/2010).

Study area

This study was conducted across the TAL-Nepal landscape (23,199 km2) that stretches along the southern lowland

areas of the country (Fig. 1). The TAL has a sub-tropical monsoonal climate and mixed deciduous vegetation ranging

from successional alluvial floodplain grasslands communities to Climax Sal (Shorea robusta) forests. This global

priority landscape [35] includes five protected areas: Suklaphanta National Park (SuNP), Bardia National Park (BNP),

Banke National Park (BaNP), Chitwan National Park (CNP), and Parsa National Park (PNP) (Fig 1). Nepal‘s

protected areas are also connected across the border with TAL region of India's 10 protected areas [18]. Based on

previous camera trap studies, adjacent CNP-PNP has the highest tiger abundance and SuNP the lowest [19].

Forest-agriculture land-use changes

Land-use changes in the past 400 years across the TAL were analyzed using the Anthrome 2.0 datasets [36].

Anthrome data sets characterize global anthropogenic transformation at the century level (i.e., every 100 years) for

the terrestrial biosphere in a discrete time frame for the years 1700–2000. We clipped the coarse data sets (~10 km

pixels size) for the landscape using ArcGIS 10.1 (Esri, Redlands, USA) and calculated the amount of area under

different land-use classes across different time frames (1700–1800–1900–2000). Thirteen land-use classes were

classified from Ellis et al. [36] into two classes (forest and agriculture-settlement) to analyze the changes from forest

to agriculture or vice versa over the time frame. This analysis was performed to identify major changes in land cover,

which was used to develop hypotheses about potential movement barriers and bottleneck events impacting Nepal‘s

tiger population over time.

Fecal DNA sampling

Putative tiger fecal (scat) samples were collected from protected areas and connecting wildlife corridors across the

TAL-Nepal. We divided the study area into 108 grid cells each measuring 15 X 15 km (225 km2, sampling unit) and

surveyed 54 grid cells (Fig.1) that had the high probability of tiger occurrence based on the Barber-Meyer et al [37]

occupancy analysis. These selected grid cells covered five protected areas and six corridors. Within each grid cell,

fecal samples were detected and collected during opportunistic field surveys. Field teams followed human trails, fire

lines, and animal trails. Fecal samples were field-identified based on their physical appearance and associated indirect

tiger signs like pugmarks and scratches [38].

A few grams from the upper surfaces of the scat were removed [39] and stored at room temperature in sterile 2-ml

vials filled with DETs buffer (dimethyl sulphoxide saline solution) [40] at1:4 volume scat-to-solution ratio. The

remaining fecal sample was left in the original location to minimize disturbance on any tiger signs. Scat samples were

transported to the Kathmandu-based laboratory of the Center for Molecular Dynamic Nepal (CMDN) for genetic

analysis.

DNA extraction, species identification and sex identification

DNA was extracted from scat samples using a commercially available QIAmp DNA Stool Kit (QIAGEN Inc.,

Germany) following the manufacturer‘s instructions. Species identification polymerase chain reaction (PCR) was

performed using tiger-specific primers (TIF/TIR) (S1 Table) [41] that amplify a 162 bp mtDNA cytochrome b

fragment. The genetically identified tiger samples were further processed for sex identification (S1,S2 Tables) [42].

Microsatellite primer selection and genotyping for individual identification

Sixteen microsatellite markers used by previous studies [41, 43-45] were evaluated. Based on amplification success

and degree of polymorphism, we amplified 10 polymorphic microsatellite loci but retain only 8 microsatellite loci for

individual identification and genetic analyses. (S3 Table).

PCR amplification was carried out in a 7 µL reaction volume containing 3.5 µL of Qiagen master-mix (Qiagen Inc.,

Germany). The PCR conditions for microsatellites amplification included an initial denaturation (95ºC for 15 min)

with a touchdown PCR step for 10 cycles (denaturation at 94ºC for 30 s, annealing initially at 62ºC and reduced by

0.5ºC in each cycle for 90 s and extension at 72ºC for 60 s). This was followed by 25 cycles of denaturation at 94ºC

for 30 s, annealing at 57ºC for 90 s and extension at 72ºC for 60 s, and a final extension at 72ºC for 10 mi. PCR

product (0.7 µL) were sized against LIZ-500 size standard in ABI 310 genetic analyzer (Applied Biosystems, USA).

Three to five PCR replicates were run per sample. Microsatellite alleles were scored in GENEMAPPER, version 4.1

(Applied Biosystems, USA). To finalize consensus genotypes, at least three identical homozygote PCR results were

assessed to determine homozygote, and each allele was observed in two independent PCR for a heterozygous

genotype [46].

Genetic analyses

PCR amplification success rates and genotype accuracy (GA) were based on results from the last two PCR runs [46].

PCR amplification success was based on the percentage of PCR successes across all the tiger-positive samples.

Genotype accuracy was calculated based on the percentage of successful PCR results that matched the finalized

consensus genotype. In addition, cumulative probabilities of identity for unrelated individuals (P(ID)) and siblings

(P(ID)sib) were estimated using Gimlet, version 1.3.3[47]. A minimum criteria of P(ID)sib [48, 49] for selecting the

minimum number of loci required for individual identification was set at 0.035. The GenAlEx program, version 6.5

[50] was used to assess genotype matching and determine the minimum number of individual tigers in the consensus

genotype data sets.

Genetic diversity was quantified by estimating observed heterozygosity (Ho) and expected heterozygosity (He) using

GenAlEx, version 6.41[50]; and allelic richness using the rarefaction method in HP-RARE, version 1.0 [51]. Global

and population-level deviations from the Hardy-Weinberg equilibrium (HWE) and linkage disequilibrium (LDE) [52]

were calculated using ARLEQUIN, version 3.5 [53] and evaluated with and without Bonferroni corrections for

multiple tests [54]. ARLEQUIN was also used to estimate FST, inbreeding coefficients (FIS) and to test the statistical

significance of the pair-wise FST values between the populations in the TAL, using 10,000 permutations [55]. This

was further complemented by the Analysis of the Molecular Variance (AMOVA), implemented within ARLEQUIN,

was used to assess the amount of variation within, and across, individuals and populations (CNP, BNP, SWR) in

TAL. DEST was used as an alternative metric for genetic distances between the populations, as it outperforms FST as an

accurate and unbiased metric of the level of differentiation between populations when the sample size is high and the

number of loci is low [56]. The harmonic mean of DEST across the loci for each population was calculated using the

web-based program SMOGD [57] with 1,000 bootstrap replicates.

To visualize genetic similarities among regions and individuals, we performed a multivariate principal coordinate

analysis (PCoA) in GenAlEx. Furthermore, an individual-based Bayesian clustering approach was implemented using

STRUCTURE (non-spatially explicit), version 2.3.4 [58] and TESS (spatially explicit), version 2.3[59], for inferring

genetic subdivision across the tiger population in the TAL. In STRUCTURE, the value k, representing the potential

number of genetic clusters (sub-populations), was allowed to vary between 1 and 5; we performed 10 independent

runs for each value of k. This analysis was run with admixture models with correlated frequencies using a burn-in of

500,000 Markov chain Monte Carlo (MCMC) steps followed by an additional 1,000,000 iterations without prior

information on the sampling sites. The optimal value of k was inferred by examining the likelihood curve, q value bar

plots, and using the Evanno method [60] implemented in the web-based program Structure Harvester [61]. The

individual membership assignments estimated in STRUCTURE were analyzed in program CLUMPP [62] with a

greedy algorithm and 10,000 random permutations for estimating the mean of the permuted matrices across

replicates. We used the stringent criterion of q> 0.8 for assigning individuals as residents to potential sub populations.

Values below 0.8 were considered representative of individuals with admixed ancestry.

The admixture models (convolution Gaussian model [BYM] and the conditional auto-regressive [CAR]) were run

using TESS [63] with 100,000 burn-in runs followed by 20,000 iterations for k = 2 to 10 with 100 replications per k.

The average of the 10% of the lowest Deviance Information Criteria (DIC) values was used for each kmax. DIC values

were taken for estimating the number of optimal kmax (genetic sub-populations) [64]. DIC values averaged over100

independent iterations were plotted against kmax and most likely value of kmax was selected by visually assessing the

point at which DIC first reached a plateau and the number of clusters to which individuals were proportionally

assigned.

We also examined isolation-by-distance (IBD) and spatial auto-correlation patterns to characterize spatial genetic

structure of tigers in TAL using GenAlEx. First, we tested whether a significant correlation existed between pair-wise

co-dominant genotypic distance and geographical distance by applying simple Mantel tests with 9,999

permutations[65]. Secondly, we used spatial auto-correlation analysis to test the spatial extent of the genetic structure

against the null hypothesis of no auto-correlation (correlation coefficient, r = 0) by generating 95% confidence

intervals (CI) for each distance class via permutation (9,999 simulations) and bootstrapping (999 repeats). We

correlated genetic distances and spatial distance matrices and generated auto-correlation matrices for each spatial

distance class ranging from 0-250 km based on the distribution of tigers in TAL. Results were visualized as

correlograms and the location of first x-intercept represents the extent of non-random spatial genetic structure.

Individuals below this threshold share a higher proportion of genes than spatially distant individuals. Within a given

correlogram, significant spatial auto-correlation was confirmed only when a positive r-value fell outside the 95% CI

(derived from the permutation test), and when the 95% CI about r (derived from bootstrapping) did not intercept the

axis of r = 0[66].

To examine whether the individuals were born in the location from which they were sampled, assignment/exclusion

tests were performed in GENECLASS, version 2.0 [67]. The Bayesian approach with re-sampling algorithm [68] was

used with 10,000 individuals at an assignment threshold (alpha, α) value of 0.01. The likelihood ratio test statistic

(Lhome/ Lmax) was applied to identify migrants. This method uses the Bayesian criteria of Rannala & Mountain (1997)

[69] along with the MCMC re-sampling method to determine the critical value of Lhome/ Lmax beyond which a sample

is treated as a migrant. We also carried out assignment tests in program STRUCTURE incorporating the geographical

sampling sites as prior information (LOCPRIOR) without changing the other parameter settings as described above.

Since we did not have prior knowledge about migration of individuals (MIGPRIOR) between potential sub-

populations, we used the default setting.

Rates of recent immigration over the last several generations among the sub-populations were estimated using

program BayesAss+, version 3.0 [70]. This Bayesian approach uses the multi-locus genotype data and relaxes the key

assumption that populations are in HWE or migration-drift equilibrium. Recent gene flow (over the past 5–7

generations, approximately 25-30 years) was assumed [27], given a 7.55 year generation time for tigers [71]. Multiple

runs (n=5) of the program BayesAss+ with 3×107 MCMC iterations and a 10

7 burn-in with different seed numbers

and delta values confirmed the final parameters and ensured their convergence. Both immigration and emigration

rates between populations were considered as contemporary migrations.

We used two approaches to test for the genetic signature of a severe demographic contraction (i.e. bottleneck) in the

tiger population across the TAL. First, we used test M-ratio [72] implemented in ARELQUIN. The M-ratio compares

the number of alleles (k) with the allelic size range (r). Presence of a bottleneck signature in a population occurs when

rare alleles are lost along with reductions in k faster than r. Low M-ratio values less than the threshold of 0.68 are

thought to represent the presence of a bottleneck signature in the population [72]. Second, we used the Cornuet and

Luikart [73] approach in program BOTTLENECK [74] for comparing the bottleneck signature in each population.

This method tests for the departure from mutation-drift equilibrium based on heterozygote excess or deficiency (Heq).

Simulations were performed under three mutation models: infinite allele model (IAM), single stepwise model (SSM),

and two-phase model (TPM). The simulation values were then compared to real data values to obtain the distribution

of Heq. For the TPM, we used the generic values of 0.95 and 0.12 for frequency of single-step mutations and variance,

respectively [74]. A Wilcoxon sign-rank test was used to detect heterozygosity excess or deficiencies across loci. We

also used a qualitative approach with the mode shift test to detect a population bottleneck. Recently bottlenecked

populations show a mode shift in the distribution of allele frequencies such that alleles with very low frequency (less

than 0.1) are less abundant than alleles that occur frequently.

Results

Landuse changes

Land-use change analysis showed a dramatic decline in forest cover and an increase in agricultural areas by 62% in

the last 300 years across the TAL-Nepal (Fig 2). The major decline in forested areas, with a 47% decrease in land

cover, occurred between the 19th and 20

th centuries. There has been a 97% increase in agriculture and settlement areas

over the past 200 years in the TAL (S1 Fig). Within the TAL, a major break in the contiguous forest landscape

occurred around 52 km west of Chitwan National Park due to a large-scale resettlement and development project

between 1900-1960s. Among these protected areas, Chitwan valley (CNP and surrounding areas) suffered the largest

and most dramatic decline in forested area. Therefore, we predicted that the CNP-PNP tiger population would be an

isolated sub-population. Based on our landscape change analysis, we expected to find a bottleneck signature within

CNP-PNP and the other hypothesized sub-populations (BNP and SuNP).

Fig 2. Land use change (forest into agriculture and settlement) analysis, in the last 400 years in the Terai Arc

Landscape using Anthrome 2.0 datasets at resolution of ~10 km pixel size [36] in ArcGIS 10.1.

Field Sampling

In the first phase of the study (2010-2012), fecal samples were collected from grid cells (n = 36) within four protected

areas (SuNP, BNP, CNP, and PNP). No fresh scats were found in Banke National Park. In the second phase (2012-

2013), putative tiger samples were only found and collected in grids (n = 5) of the two known corridors (Khata and

Basanta). No fresh tiger scats were found in corridors: Kamdi, Laljhadi, Barandabhar, and Bramahadev.

Genetic analysis of tiger samples

Of 770 putative tiger scat samples collected from four protected areas (SuNP, BNP, CNP, and PNP) and two

corridors (Khata and Basanta) within the TAL, a total of 412 were confirmed to be tiger scat (Fig 3) (Table1). Of

these, sex was genetically identified in 353 scat samples, among which 255 samples came from males and rest (98)

were from females (Fig 4) (Table 1). Of the ten microsatellite loci that were amplified, eight were retained for

individual identification and genetic analysis based on high PCR amplification success (84%), genotyping accuracy

(82%) (Table 2), and polymorphism. Two Loci (FCA205 and PttA2) were removed from the analysis due to poor

amplification success or monomorphism. For the remaining 8 loci, the cumulative P(ID) and P(ID)sib were estimated to

be 1.5E-06 and 3.2E-03 (<0.01) respectively, and 3.1 E-02 for our 6 least powerful loci. We obtained a consensus

genotype at 6-8 loci for 212 samples (51% genotyping success) and identified 78 individual tigers (male=49,

female=27, unknown sex=2). Only four scat samples collected outside of protected areas were of tiger (Table 1).

Unfortunately, we were unable to identify the number of individuals from these samples due to poor DNA quality.

Only one tiger sample from the PNP collected 15 km east of the CNP was successfully genotyped and added to the

CNP dataset.

Fig 3. Tiger species identification using mtDNA PCR assay. Tiger positive samples yielded 162bp PCR

fragments.

Fig 4. Sex identification of tiger species positive samples using Amelogenin gene PCR assay. Females ha ve

single band at 214 bp and males have bands at 194 bp and 214 bp.

Table 1. Summary of scat DNA analysis success rates for species and sex identification, and overall

microsatellite genotyping based on putative tiger scat samples (n = 770) collected in the Terai Arc Landscape,

Nepal. Species identification was based on the total number of samples processed; sex identification was based

on the total number of tiger positive samples; and genotyping success was based on the number of positive

samples for species and sex identification. M, male; F, female.

Study Areas

Total Samples

Processed

Tiger Species

Identification

% (# samples)

Sex Identification

% (# samples; M, F)

Genotyping

Success

% (# samples)

Chitwan National Park 420 61 (257) 93 (239; 185, 54) 56(145)

Bardia National Park 116 67 (78) 85 (66; 46, 20) 55(43)

Suklaphanta National Park 79 83 (66) 59 (39; 19, 20) 35(23)

Parsa National Park 85 8 (7) 85 (6; 4, 2) 14(1)

Corridors 70 6 (4) 75 (3; 1, 2) 0(0)

All Areas 770 54 (412) 86 (353; 255, 98) 51(212)

Table 2. Summary of PCR amplification success and genotyping accuracy (GA) for 8 microsatellite loci for all processed tiger samples (n = 401)

detected across the three protected areas: Chitwan National Park, Bardia National Park, and Suklaphanta National Park across the Terai Arc

Landscape, Nepal.

Locus

All Samples (n=401) Chitwan NP (n=257) Bardia NP (n=78) Suklaphanta NP (n=66)

PCR GA PCR GA PCR GA PCR GA

FCA391 75.7 88.6 87.6 89.5 57.5 89.9 81.9 86.4

PttD5 88.4 88.8 91.7 92.0 83.3 83.8 90.3 90.6

FCA232 87.5 77.0 82.8 80.2 82.5 65.6 97.2 85.3

FCA304 94.0 87.5 93.9 90.2 90.8 79.4 97.2 92.7

FCA043 93.6 80.3 95.2 89.0 92.5 80.6 93.1 71.2

F53 65.8 84.6 69.7 81.9 33.3 79.4 94.4 92.5

F85 73.9 73.4 75.5 85.1 89.2 57.1 56.9 78.0

FCA441 94.8 74.7 94.3 85.8 91.7 71.3 98.6 67.2

Mean 84.2 81.9 86.3 86.7 77.6 75.9 88.7 83.0

SD 11.0 6.3 9.5 4.2 21.2 10.6 13.9 9.8

PCR, % polymerase chain reaction amplification success; GA, % genotyping accuracy; NP: National Park; SD: Standard deviation

Equilibrium analyses and genetic diversity

Based on the population-level analysis, four loci in CNP, two in BNP, and three in SuNP deviated significantly from

Hardy-Weinberg Equilibrium (HWE) at P < 0.05 (Table 3). No locus was consistently out of the HWE across all

sampling sites. After Bonferroni corrections, one locus (F85) in CNP and one locus (F53) in SuNP remained

significantly out of the HWE. Significant linkage disequilibrium after sequential Bonferroni corrections (P≤ 5.95E-

04) was detected among three pairs of loci with no apparent pattern (FCA232-FCA043 in CNP, FCA304-F53 and

FCA391-FCA232 in SuNP). Overall average genetic diversity of the TAL tiger population was as follows: observed

heterozygosity of 0.54, expected heterozygosity of 0.61, mean number of alleles per locus of 6.0, and mean allelic

richness of 3.51. Levels of genetic variability varied among the protected areas, with the highest genetic diversity

found in CNP and the lowest in SuNP (Table 3). The average local inbreeding coefficient was high (FIS = 0.16) in the

smaller SuNP population relative to other sites in the TAL (Table 3), which is consistent with the lower diversity

observed for SuNP population.

Table 3. Genetic diversity estimates across 8 microsatellite loci for tigers studied in three protected areas across the Terai Arc Landscape, Nepal: Chitwan

National Park, Bardia National Park, and Suklaphanta National Park.

n, sample size; NA, number of alleles; AR, allelic richness using the rarefaction method; Ho, observed heterozygosity; He, expected heterozygosity; PHW, P values for exact

tests of Hardy-Weinberg equilibrium (level of significance, α= 0.05); FIS, inbreeding coefficient. * represents locus out of HWE after Bonferonni correction at P=0.002;

SE, standard error.

Locus

Chitwan National Park (n=37) Bardia National Park (n=25) Suklaphanta National Park (n=16)

NA AR Ho HE PHW FIS NA AR Ho HE PHW FIS NA AR Ho HE PHW FIS

FCA391 4.00 3.96 0.78 0.69 0.67 -0.11 3.00 2.95 0.26 0.42 0.00 0.39 3.00 2.78 0.31 0.51 0.10 0.41

PttD5 4.00 2.66 0.32 0.30 0.01 -0.07 4.00 3.75 0.54 0.53 0.93 -0.01 3.00 2.99 0.63 0.54 0.05 -0.12

FCA232 3.00 2.89 0.74 0.55 0.04 -0.35 5.00 3.84 0.39 0.40 0.79 0.03 3.00 2.56 0.13 0.12 1.00 -0.02

FCA304 5.00 3.62 0.65 0.51 0.11 -0.27 5.00 3.75 0.52 0.44 0.61 -0.16 2.00 2.00 0.19 0.48 0.03 0.63

FCA043 4.00 3.87 0.51 0.61 0.01 0.17 5.00 4.28 0.57 0.53 0.56 -0.04 4.00 3.77 0.38 0.54 0.07 0.33

F53 5.00 4.26 0.67 0.67 0.42 0.02 5.00 4.91 0.56 0.75 0.03 0.29 5.00 5.00 0.67 0.73 0.00* 0.13

F85 6.00 4.83 0.42 0.64 0.00* 0.36 4.00 3.50 0.63 0.61 0.90 0.00 3.00 3.00 0.73 0.64 0.91 -0.12

FCA441 4.00 3.30 0.57 0.58 0.89 0.03 4.00 3.99 0.92 0.72 0.54 -0.26 3.00 3.00 0.64 0.62 0.89 0.00

Average 4.00 3.67 0.58 0.57

-0.03 4.00 3.87 0.57 0.55

0.03 3.00 3.14 0.46 0.52

0.16

SE 0.32 0.72 0.06 0.04

0.08 0.26 0.57 0.04 0.05

0.07 0.31 0.90 0.08 0.06

0.09

Genetic structure

The mean global FST value (level of differentiation) for the TAL was found to be moderate (FST = 0.14

± 0.07). Results of AMOVA showed that genetic variation among the sites was 13.7%, while the

residual variation among individuals within the sites was 86.3% (S4 Table) indicating a low level of

differentiation among the populations for the target species. The pair-wise FST and DEST values

(between protected areas) were found to be low to moderate and statistically significant (Table 4, P <

0.05). Comparatively low FST and DEST were found between the BNP and SuNP populations. In

contrast, levels of genetic differentiation were high (FST = 0.21) for SuNP and CNP populations (Table

4).

Table 4. Pair-wise measures of the level of differentiation of tiger sub-populations in the Terai

Arc Landscape, Nepal based on FST[55] and DEST (in parentheses) [56] (below the diagonal). Pair

wise geographical distance (in km) between core population (above diagonal).

Population CNP BNP SuNP

CNP ____ 314 450

BNP 0.08 (0.07) ____ 136

SuNP 0.21 (0.21) 0.14 (0.12) ____

All pair-wise FST values were significant (P<0.05) based on 10,000 permutations.

The principal coordinate analysis performed in GenAlEx 6.5 showed that individual tigers formed

three not so distinct overlapping clusters matching their sampling localities (CNP, BNP, SuNP) (Fig

5). Results of the Bayesian clustering analysis in STRUCTURE indicated three discrete, core

populations across the TAL with some genetic admixture. The models showed highest statistical

support at k = 4 based on the Ln P(k), and k = 3 based on delta k method. Four clusters showed a high

standard deviation relative to k = 3 (S2 Fig) and sub-structure within CNP (S3 Fig). Hence, k = 3 was

interpreted as the most likely value for the analysis, and it aligned with the prior knowledge of the

spatial distribution of tiger demographic populations and land-use analysis (Fig 6). At k=2, a major

split was detected between CNP and the other two protected areas (BNP and SuNP, S3 Fig), which is

consistent with the land use data and spatial separation of these sites (Fig 1 and 2).

Fig 5. Principal Coordinate Analysis of tiger genotypes obtained from “♦” Chitwan National

Park, “○” Bardia National Park, and “▲” Suklaphanta National Park, in the Terai Arc

Landscape, Ne pal, assessed through program GenAlEx.

Fig 6. Map of the Terai Arc Landscape, Nepal. (a) Protected areas (starting from left:

Suklaphanta National Park, SuNP; Bardia National Park, BNP; Chitwan National Park, CNP;

along with the spatial location of identified tiger-positive samples (black dots). (b) Pie charts

showing the percentage of ancestry assigned to other identified genetic clusters in the

populations (orange), and the resident population (blue). (c) STRUCTURE (non-spatially

explicit) bar plot with each bar representing an individual tiger (n = 78) in three populations

across the Terai Arc Landscape revealing three (k = 3) admixed sub-populations (represented

by 3 different colors) along with five migrants (marked as “*’) identified across the population.

(d) Bar plot showing three identified sub-populations analyzed in spatially-explicit assignment

program TESS.

The inferred degree of admixture using STRUCTURE within populations was found to be variable

across the three clusters (CNP, BNP, and SuNP) with the central cluster (BNP) showing the highest

degree of admixture: 1st cluster (91% CNP, 7% BNP, and 2% SuNP); 2

ndcluster (13% CNP, 72%

BNP, and 15% SuNP); and third cluster (2% CNP, 7% BNP, and 91% SuNP) (Fig 6). Within inferred

clusters, sites (protected areas) are spatially separated from each other (1st- 2

nd clusters: 314 km; 2

nd

and 3rd

clusters: 136 km; and 1st and 3

rd cluster: 450 km), with variable forest connectivity between

them (S5Table, Fig 7).

Fig 7.Contemporary gene flow patterns for tigers inferred across the Terai Arc Landscape,

Nepal, based on migration rates (Mc) estimated in BayesAss+[70]. Dashed lines in the center

indicate direction of migration and line thickness represents the magnitude of estimates along

with the migration rates. Figures within parentheses represent 95% CI for migration rates. Size

of the circle represents the estimated size of breeding population. Broken lines around the

periphery represent the spatial distances between the populations.

Results from the non-spatially explicit clustering analysis in STRUCTURE were corroborated with

the spatially explicit analyses in TESS. Based on the DIC model selection criteria, graphs tended to

plateau at k = 3, and the standard deviation increased with an increasing value of kmax (S4 Fig). The

hard-clustering algorithm for individual membership with BYM admixture models did not change

with kmax≥ 3, while in the CAR admixture model, there were inconsistent results with kmax> 3. Overall,

k = 3 was the best supported model, inferred number of genetic clusters for the tiger populations

across the TAL, thereby identifying CNP (1), BNP (2), and SuNP (3) as three distinct sub-populations

with some evidence for sub-structure within CNP.

There was significant correlation (r = 0.48, P= 0.0001) between geographical distance and genetic

distance (Fig 8), supporting the hypothesis of isolation-by-distance for tigers in the TAL. The auto-

correlogram for all tigers showed significantly positive autocorrelation within 75 km distance class

(25 km, r = 0.050, P = 0.0001; 50 km, r = 0.094, P = 0.0001; 75 km, r = 0.076, P = 0.02). An x-

intercept of r at ~ between 75-100 km (S5 Fig), empirically shows presence of nonrandom spatial

structuring and genetic association among individuals at distances <94 km.

Fig 8. Isolation-by-distance patterns for tigers in Terai Arc Landscape, Nepal assessed by

plotting pairwise codominant genotypic distance versus pairwise geographic distances (km)

including statistical significance using simple Mantel tests in GenAlEx, version 6.5. Each point

(diamond) represents a pair-wise comparison among individual tigers.

Detection of migrants

We identified a total of seven migrants across the TAL using our criteria. STRUCTURE identified

five migrants, while Geneclass2 identified four migrants and BayesAss+ identified six (Table 5).

Three migrants were identified by all three methods – one from BNP to CNP (male), and two from

CNP to BNP (male). Two female tigers were classified as migrants by both STRUCTURE and

BayesAss+ moving from SuNP to BNP. Geneclass2 and BayesAss+ identified two additional

migration events (both male tigers) moving from the BNP to CNP and the SuNP to BNP.

STRUCTURE results suggested that both male tigers had admixed genetic ancestry (Table 5). No

migrants were detected between the CNP and SuNP sub-populations.

Table 5. First-generation migrants between the 3 main core tiger populations in the Terai Arc

Landscape-Nepal detected using programs STRUCTURE, Geneclass2 and BayesAss+; P value

from Geneclass2.

Tiger ID*

Sampling

Location

P Value

STRUCTURE

(estimated ancestry

from sampling

locations)

Assigned

Population

Sex

CHIT001(S,G,B)

CNP 0.0001 0.03 BNP Male

CHIT011(G)

CNP 0.009 0.94 BNP Male

BRD001(S,G,B)

BNP 0.002 0.14 CNP Male

BRD002(S, G,B)

BNP 0.007 0.21 CNP Male

BRD023(B)

BNP 0.04 0.62 SuNP Male

BRD024(S, B)

BNP 0.53 0.23 SuNP Female

BRD025(S,B)

BNP 0.53 0.24 SuNP Female

*Superscript caption with ID number indicates that individuals were identified as migrants either by

STRUCTURE (S) or Geneclass2 (G) or BayesAss+ (B).

The BayesAss+ analysis showed symmetric migration between pairs of populations (Fig 7), except for

SuNP. A high immigration rate was determined from the SuNP into the BNP sub-population (mc =

0.13). The BNP sub-population appeared to be receiving the migrants from both CNP and SuNP, with

a total migration rate of 0.17 (Fig 7). The 95% confidence intervals overlapped for migration rates for

the pair of populations (SuNP-CNP; BNP-CNP) in both directions, suggesting roughly symmetric bi-

directional gene flow between sub-populations. This did not hold for gene flow between BNP and

SuNP, which appeared to be uni-directional from SuNP to BNP. The net emigration rate was highest

for SuNP, as it contributed the most migrants to the other protected areas (Fig 7). The net emigration

rate for BNP and CNP were negative, suggesting the site received more migrants than it contributed to

other populations.

Detection of bottlenecks

The average M-ratio describing potential bottlenecks for all sites was 0.29 (SE = 0.07), which is

below the threshold value of 0.68. This suggests that the tiger population suffered a bottleneck event

that caused a severe decline in the population size in the recent past. These results were supported

among all sub-populations. The Bottleneck software detected heterozygote excess using the sign test,

with 5 to 6 loci, depending upon the mutation models. However, results of the two-tailed Wilcoxon

signed-rank test showed that the signature of a bottleneck event was significant only for the CNP

population under the Infinite Allele Mutational (IAM) model (P = 0.03, S6 Table). The mode-shift

test showed the normal ―L‖-shaped allele distribution in BNP. The test only showed the presence of a

large proportion of alleles at low frequencies, indicating a genetic bottleneck between CNP and SuNP,

but not CNP and BNP.

Discussion

Effect of landuse change

The historical land-use change analysis was consistent with the present structuring of animals on the

landscape into sub-populations [27, 75, 76]. Tigers used to roam the vast expense of Terai forests in

Nepal and India. Beginning in the mid-1840s, core tiger habitat was protected by the Rana rulers for

their exclusive royal hunting, thus discouraging people from settling and conducting agriculture.

Tigers were persecuted in large numbers during organized royal hunts. However, the population

recovery of tigers was relatively quick following this time, likely due to immigration from adjacent

forest areas. Rampant malaria also hindered people from clearing forest and settling in the Terai

region [77].

By the late 20th century however, extensive land clearing had occurred and the level of wildlife

hunting was high, fragmenting tiger habitat, reducing abundance of tiger prey species, and contracting

tiger range to the Chitwan Valley in central Nepal [75, 78], the BNP and SuNP in the west, and to a

few large blocks of forest in the east. High human population density and extensive deforestation for

agricultural practices [79] led to inadequate cover and low prey availability and tigers became

extirpated in the eastern Terai region by the 1970s [80].With the eradication of malaria and the

government policy of extending settlements along the border with India, large swaths of forest

landscape were disturbed [81]. Taken together, these events resulted in the loss and fragmentation of

the forest biomes in the TAL, likely causing the decline in tiger population size and genetic diversity,

and the subsequent sub-structuring of the tiger population in Nepal, perhaps through population

bottlenecks [36].

High rates of anthropogenic transformation of landscapes [36] play a major role in the extinction of

wild mammal populations [82]. Intense forest fragmentation has imposed similar effects upon jaguars

in Brazil [83] and Amur leopards in the Russian Far East [84], causing significant loss of genetic

variation. In contrast, tigers [27, 85] and leopards [86] in India have high levels of genetic variation,

and no genetic bottlenecks have been detected despite habitat fragmentation.

Sampling Strategy

Our sampling strategy was to collect as many as possible from the potential habitat (protected areas,

surrounding buffer zones forests, and corridors) along TAL. We screened the duplication of samples

at two stages: at field, we only collected the fresh scat samples at each encounter and avoided

collection of the same scats samples by putting in the red marker on the scat (dried) itself and/or

placed a twig into soil at the site of fresh samples. At the analysis phase, each of the duplicate samples

(if scored) were taken as recapture.

Genetic variation in TAL

Tigers in the TAL displayed moderate levels of genetic variation (He = 0.61) across the landscape and

similar levels across sub-populations (He ≈ 0.57), perhaps due to moderate population size and/or

gene flow resulting from potential connectivity among tiger-bearing protected areas across the TAL in

Nepal (n=5) and India (n=10).Our estimates of genetic variation were lower in comparison to genetic

diversity estimates for Bengal tigers overall (He = 0.72) [87] or the Indian Subcontinent (He = 0.70)

[22] using microsatellite loci. However, the estimates were higher than in other sub-species of tigers

(average He = 0.53, Sumatran, Indochinese, Malayan and Siberian) [87]. Landscape-wide genetic

variation in tigers across the major tiger conservation landscapes in India range from as low as He =

0.58 in the southeastern Ghats [88], He = 0.67 in the northeast landscape [89], He = 0.76 in the

Western Ghats [24], and as high as He = 0.81 in the central Indian landscape [27]. However, care

should be taken in interpretation of heterozygosity because direct comparisons of diversity are not

possible since these studies employed different numbers and combinations of microsatellite loci.

The average inbreeding coefficient across the sub-populations in the TAL-Nepal suggests weak

inbreeding that was statistically non-significant (P= 0.42). At the sub-population level, SuNP showed

a weak sign of local inbreeding (FIS = 0.16), which approached statistical significance (P= 0.06). This

suggests the importance of connectivity in averting inbreeding depression in the small SuNP sub-

population[90]. Spatially, SuNP is surrounded by human settlements, but retains dispersal potential

through the northern and southern sections of the reserve. However, we did not detect any migrants

into this population, but rather detected three migrants leaving this subpopulation. Additionally, the

creation of the Pilibhit Tiger Reserve (India) to the south strengthens the possibility of future

connectivity via tiger dispersal through the Lagga Bagga Forest in India (Dr A.J.T Johnsingh:

personal communication). Currently, the net migration estimates suggest that more tigers are leaving

SuNP than are immigrating (Fig 7).

Genetic structure

Both population- and individual-based tests for assessing the level of genetic sub-division revealed

moderate levels of differentiation across the landscape. This is in concordant with AMOVA and

fixation index (Fst) results where populations (sites) observed moderate differentiation showing three

genetic clusters confirming that total population in Terai Arc may not be total panmixia. Our Bayesian

clustering results support three distinct genetic clusters (sub-populations) within the TAL-Nepal,

representing the three tiger bearing protected areas (CNP, SuNP, BNP) confirming our a priori

hypothesis. Our field data suggested demographic contiguity in and around these protected areas [30].

However, with the absence of tiger samples from the Indian side of the TAL, there is a lack of clarity

in explaining the overall genetic structure of tigers across the entire landscape; tigers could travel

through the contiguous protected areas of India into Nepal. Joint camera trap surveys conducted

across the transboundary protected areas revealed 10 common tigers dispersing between Nepal and

India [91]. We did detect a strong isolation-by-distance effect for individual tigers across the TAL.

Spatial auto-correlation analysis detected genetic spatial structuring at geographic scale of 75-100 km

across all samples indicating high genetic association among individuals even at such broad distances,

and gene flow for the tigers might be possible through the landscape. The Bayesian clustering analysis

in STRUCTURE and migration analysis in BayesAss+, both assessing contemporary gene flow levels

[92, 93], showed higher connectivity between SuNP and BNP than between BNP and CNP. In

contrast FST/DEST showed higher connectivity between CNP and BNP, which are farther apart, than

between SuNP and BNP, which are closer together.

Gene flow and detection of migrants

The results of the assignment tests offered evidence of contemporary gene flow between tiger sub-

populations suggesting that tigers can, and do, disperse across the TAL-Nepal. We detected 7

migrants (5 males, 2 females) moving between the sub-populations. Tigers are known to disperse long

distances (~200 km) based on long-term camera-trap data [94]. Furthermore, dispersal has been

recorded as far as 600 km in the central Indian landscapes [25]. Consequently, the detection of tiger

dispersal (first-generation migrants) between protected areas, which range from 314 km (CNP and

BNP) to 136 km (BNP and SuNP) apart, is expected if the landscape provides stepping stone habitats

to allow tiger movement throughout the landscape.

Results of the BayesAss+ analysis showed high estimates of recent migration of tigers among the sub-

populations. High net immigration rates were detected in BNP. In the medium and large sub-

populations (BNP and CNP), we found evidence of one to two male tigers dispersing between the

populations. There was also evidence of one male and two females moving from SuNP to BNP. If

these tigers breed, this will likely avert the detrimental effects of inbreeding depression, thus

flattening the slopes of extinctions curves accordingly [90]. Results from the STRUCTURE analyses

(Table 5) detected 3-7 individuals with mixed ancestry suggesting that migrants have successfully

reproduced in earlier generations.

The SuNP population, while small, holds the highest density of tigers, likely due to high prey density

[19]. This could help explain why SuNP is the source of many emigrants. However, it could be argued

that, due to the small size of the reserve and the high tiger density, there might not be space for

dispersing male tigers to establish territories within the reserve. Hence, it may not be possible to

receive migrants from surrounding areas. Lower levels of genetic variation, and the slight but weak

evidence of inbreeding relative to other protected areas in the landscape, suggests the need to increase

migration of tigers into the SuNP population to avert inbreeding depression and increase genetic

variation of the subpopulation. For example, leopards in the Russian Far East suffered significant loss

in genetic variation due to lack of connectivity to a source population and continue to suffer loss of

genetic variation [84]. The designation of 727 km2 of the Pilibhit Forest Division as a tiger reserve by

the Indian government is an important step towards increasing connectivity between the western TAL

in India and SuNP in Nepal (Dr. Dipankar Gosh WWF India: personal communication).

Population bottlenecks

While both analysis techniques revealed population bottlenecks in the TAL-Nepal, there was a

disagreement in the resulting number of detected population bottlenecks. The M-ratio test revealed a

population bottleneck in all populations, in contrast to the heterozygous excess test showing a

bottleneck only in the CNP population. Both tests have been found to be effective at detecting

bottlenecks, but each works under different assumptions. Peery et al. [95] suggested that despite

correctly assuming the mutation models (IAM, SSM or TPM), statistical power to detect a bottleneck

with the two methods might depend upon the pre-bottleneck genetic variation. Heterozygosity may be

less powerful than the M-ratio test when pre-bottleneck genetic diversity is high [95]. Alternatively,

the heterozygosity excess test may work best when the pre-bottleneck population is smaller or when

the bottleneck is milder and more recent[96]. Either way, there is evidence that at least one bottleneck

occurred in the TAL-Nepal.

Conclusions

We provide evidence of three genetically admixed sub-populations across the TAL-Nepal based on

spatial (TESS) and non-spatial (STRUCTURE) Bayesian clustering techniques, suggesting that tigers

have been able to move between the populations and breed, at least in the recent past. Contemporary

gene flow measures of tigers were estimated in the TAL based on both likelihood and Bayesian

approaches. Improved connectivity between the protected areas, facilitated by male and female tiger

dispersal, appears to be able to avert the negative consequences of inbreeding depression following

bottleneck events. Thus, there is a need to maintain connectivity throughout the TAL-Nepal landscape

and beyond. We did not find migrants from CNP into SuNP or vice versa in any of the migrant

detection tests. However, the dispersal among SuNP and CNP is more likely to be a stepping-stone

process [97].

In recent times, connectivity in the Nepali landscape has been improved by the protection of large

forest blocks after the nationalization of forests in the 1960s [77], improved governance, and

management of forest resources [98]. The launch of the community forest program in the ―Terai

forest‖ [99] in the early 1980s has been successful in building more forest habitat at forest edges as

well as buffers for the large forest blocks with the goal to increase tiger dispersal across the landscape.

The Government of Nepal (GoN) has also taken positive steps in restoring connectivity across the

TAL, including the adoption of a successful, community-based forest management approach [29]. The

GoN‘s declaration of all the identified forest corridor as a ―protected forest‖ status was also a

milestone [100]. Our results indicate that the landscape is currently functional with respect to the

dispersal of tigers among the protected areas, but there is evidence of genetic structure, indicating that

sub-populations exist and gene flow is limited between some protected areas. In the face of human

population growth, economic development post-insurgency, political unrest, and developmental road

projects, genetic connectivity seems likely to erode [101]. Consequently, gene flow of tigers across

the landscape will be impeded, thus lowering their persistence in the long run [90]. Therefore, it is

essential to secure tiger core areas and functional forest corridors between them to maintain and

enhance gene flow across the landscape, thereby ensuring that tiger populations exist for generations

to come. Securing tiger existence across the TAL-Nepal will require concerted stakeholder

cooperation, careful planning, and the prevention of detrimental development activities.

Acknowledgments

We would like to thank the Department of National Parks and Wildlife Conservation for permission to

conduct the first genetic study of tigers in Nepal. We thank all the field coordinators (chief wardens)

for their coordination of fieldwork and all the field assistants and volunteers for collection of samples.

We would like to thank National Trust for Nature Conservation for their support during the field

work. Thanks to Drs: Eric Hallerman, Jess Jones, and Jamie Roberts for their guidance in analyzing

the genetic data. Finally, thanks to Dr. Jennifer Adams and University of Idaho Laboratory for

Ecological, Evolutionary and Conservation Genetics for laboratory assistance.

Conflict of interest

The authors declare no conflicts of interest

Data availability

All data related to the study is included in the manuscript and the final Nepal Tiger Genome

Project final report is available at

https://figshare.com/articles/Assessment_of_genetic_diversity_population_structure_and_gene_

flow_of_tigers_Panthera_tigris_tigris_across_Nepal_s_Terai_Arc_Landscape/5896063

References

1. Lacy RC. Importance of genetic variation to the viability of mammalian populations. Journal of Mammalogy. 1997:320-35. 2. Burkey TV. Extinction in nature reserves: the effect of fragmentation and the importance of migration between reserve fragments. Oikos. 1989:75-81. 3. Soulé ME, Alberts AC, Bolger DT. The effects of habitat fragmentation on chaparral plants and vertebrates. Oikos. 1992:39-47. 4. Avise JC. Introduction: the scope of conservation genetics. Conservation genetics: Springer; 1996. p. 1-9.

5. Mills LS, Allendorf FW. The one‐migrant‐per‐generation rule in conservation and management. Conservation Biology. 1996;10(6):1509-18. 6. Baudry J, Burel F, editors. Dispersal, movement, connectivity and land use processes. Key Concepts in Landscape Ecology, Proceedings of the 1998 European Congress of the International Association for Landscape Ecology; 1998. 7. Wright S. The roles of mutation, inbreeding, crossbreeding, and selection in evolution: na; 1932. 8. Harrison S, Hastings A. Genetic and evolutionary consequences of metapopulation structure. Trends in Ecology & Evolution. 1996;11(4):180-3. 9. Crnokrak P, Roff DA. Inbreeding depression in the wild. Heredity. 1999;83(3):260-70. 10. Keller LF, Waller DM. Inbreeding effects in wild populations. Trends in Ecology & Evolution. 2002;17(5):230-41. 11. Carbone C, Gittleman JL. A common rule for the scaling of carnivore density. Science. 2002;295(5563):2273-6. 12. Caughley G. Directions in conservation biology. Journal of animal ecology. 1994:215-44. 13. Frankham R, Briscoe DA, Ballou JD. Introduction to conservation genetics: Cambridge university press; 2002. 14. Sanderson E, Forrest J, Loucks C, Dinerstein E, Ginsberg J, Seidensticker J, et al. Setting Priorities for the Conservation and Recovery of Wild Tigers: 2005–2015. The Technical Assessment. Washington, DC: WCS, WWF, Smithsonian and NFWF-STF. 2006. 15. GTRP. Global Tiger Recovery Program. 2010. 16. Seidensticker J. Saving wild tigers: a case study in biodiversity loss and challenges to be met for recovery beyond 2010. Integrative Zoology. 2010;5(4):285-99. 17. Wikramanayake E, McKNIGHT M, Dinerstein E, Joshi A, Gurung B, Smith D. Designing a conservation landscape

for tigers in human‐dominated environments. Conservation Biology. 2004;18(3):839-44. 18. Jhala Y, Qureshi Q, Gopal R, Sinha P. Status of the Tigers, Co-predators, and Prey in India, 2010. New Delhi,: Wildlife Institute of India Dehradun; 2011. p. 302. 19. Dhakal M, Thapa M, Jnawali S. Status of tigers and prey in Nepal Department of National Park and Wildlife Conservation. Nepal, Kathmandu Google Scholar. 2014. 20. Wilson EO, Willis EO. Applied biogeography. In: Cody ML, Diamond JM, editors. Ecology and evolution of communities. 522: Harvard University Press; 1975. p. 534. 21. Beier P, Noss RF. Do habitat corridors provide connectivity? Conservation Biology. 1998;12(6):1241-52. 22. Mondol S, Karanth KU, Ramakrishnan U. Why the Indian subcontinent holds the key to global tiger recovery. PLoS Genet. 2009b;5(8):e1000585. 23. Sharma R, Stuckas H, Bhaskar R, Rajput S, Khan I, Goyal SP, et al. mtDNA indicates profound population structure in Indian tiger (Panthera tigris tigris). Conservation genetics. 2009;10(4):909-14. 24. Reddy PA, Gour DS, Bhavanishankar M, Jaggi K, Hussain SM, Harika K, et al. Genetic Evidence of Tiger Population Structure and Migration within an Isolated and Fragmented Landscape in Northwest India. PLoS ONE. 2012;7(1):e29827. doi: 10.1371/journal.pone.0029827. 25. Joshi A, Vaidyanathan S, Mondol S, Edgaonkar A, Ramakrishnan U. Connectivity of tiger (Panthera tigris) populations in the human-influenced forest mosaic of central India. PloS one. 2013;8(11):e77980. 26. Sharma S, Dutta T, Maldonado JE, Wood TC, Panwar HS, Seidensticker J. Forest corridors maintain historical gene flow in a tiger metapopulation in the highlands of central India. Proceedings of the Royal Society of London B: Biological Sciences. 2013a;280(1767):20131506. 27. Sharma S, Dutta T, Maldonado JE, Wood TC, Panwar HS, Seidensticker J. Spatial genetic analysis reveals high connectivity of tiger (Panthera tigris) populations in the Satpura–Maikal landscape of Central India. Ecology and evolution. 2013;3(1):48-60. 28. Smith JLD, Ahern SC, McDougal C. Landscape analysis of tiger distribution and habitat quality in Nepal. Conservation Biology. 1998;12(6):1338-46. 29. Wikramanayake E, Manandhar A, Bajimaya S, Nepal S, Thapa G, Thapa K. The Terai Arc Landscape: a tiger conservation success story in a human-dominated landscape. Tigers of the world: the science, politics, and conservation of Panthera tigris, 2nd ed Elsevier/Academic Press, Oxford, UK. 2010:161-72.

30. Barber‐Meyer S, Jnawali S, Karki J, Khanal P, Lohani S, Long B, et al. Influence of prey depletion and human

disturbance on tiger occupancy in Nepal. Journal of Zoology. 2013;289(1):10-8. 31. Smith JLD. The role of dispersal in structuring the Chitwan tiger population. Behaviour. 1993;124(3):165-95.

32. Wikramanayake E, Dinerstein E, Seidensticker J, Lumpkin S, Pandav B, Shrestha M, et al. A landscape‐based conservation strategy to double the wild tiger population. Conservation Letters. 2011;4(3):219-27.

33. Acharya K, Dangi R. Case studies on measuring and assessing forest degradation: forest degradation in Nepal: review of data and methods. 2009. 34. Thapa K, Wikramanayake E, Malla S, Acharya KP, Lamichhane BR, Subedi N, et al. Tigers in the Terai: Strong evidence for meta-population dynamics contributing to tiger recovery and conservation in the Terai Arc Landscape. Plos One. In Press.

35. Wikramanayake ED, Dinerstein E, Robinson JG, Karanth U, Rabinowitz A, Olson D, et al. An Ecology‐Based Method for Defining Priorities for Large Mammal Conservation: The Tiger as Case Study. Conservation Biology. 1998;12(4):865-78. 36. Ellis EC, Klein Goldewijk K, Siebert S, Lightman D, Ramankutty N. Anthropogenic transformation of the biomes, 1700 to 2000. Global Ecology and Biogeography. 2010;19(5):589-606. 37. Barber-Meyer SM, Jnawali SR, Karki JB, Khanal P, Lohani S, Long B, et al. Influence of prey depletion and human disturbance on tiger occupancy in Nepal. Journal of Zoology. 2013;289(1):10-8. doi: 10.1111/j.1469-7998.2012.00956.x. 38. Karanth KU, Sunquist ME. Population structure, density and biomass of large herbivores in the tropical forests of Nagarahole, India. Journal of Tropical Ecology. 1992;8(01):21-35. 39. Wultsch C, Waits LP, Hallerman EM, Kelly MJ. Optimizing collection methods for noninvasive genetic sampling of Neotropical felids. Wildlife Society Bulletin. 2015;39(2):403-12. 40. Frantzen M, Silk J, Ferguson J, Wayne R, Kohn M. Empirical evaluation of preservation methods for faecal DNA. Molecular Ecology. 1998;7(10):1423-8. 41. Bhagavatula J, Singh L. Genotyping faecal samples of Bengal tiger Panthera tigris tigris for population estimation: a pilot study. BMC genetics. 2006;7(1):48. 42. Pilgrim K, McKelvey K, Riddle A, Schwartz M. Felid sex identification based on noninvasive genetic samples. Molecular Ecology Resources. 2005;5(1):60-1. 43. Mondol S, Karanth KU, Kumar NS, Gopalaswamy AM, Andheria A, Ramakrishnan U. Evaluation of non-invasive genetic sampling methods for estimating tiger population size. Biological Conservation. 2009a;142(10):2350-60. 44. Menotti-Raymond M, David VA, Lyons LA, Schäffer AA, Tomlin JF, Hutton MK, et al. A genetic linkage map of microsatellites in the domestic cat (Felis catus). Genomics. 1999;57(1):9-23. 45. Mishra S, Sharma R, Singh SK, Munjal AK, Goyal SP. A comparative study of the use of tiger-specific and heterologous microsatellite markers for population genetic studies of the Bengal tiger (Panthera tigris tigris). African Journal of Biotechnology. 2014;13(8):936-43. 46. Wultsch C, Waits LP, Kelly MJ. Noninvasive individual and species identification of jaguars (Panthera onca), pumas

(Puma concolor) and ocelots (Leopardus pardalis) in Belize, Central America using cross‐species microsatellites and faecal

DNA. Molecular ecology resources. 2014;14(6):1171-82. 47. Valière N. GIMLET: a computer program for analysing genetic individual identification data. Molecular Ecology Notes. 2002;2(3):377-9. 48. Waits LP, Luikart G, Taberlet P. Estimating the probability of identity among genotypes in natural populations: cautions and guidelines. Molecular ecology. 2001;10(1):249-56. 49. Mills LS, Citta JJ, Lair KP, Schwartz MK, Tallmon DA. Estimating animal abundance using noninvasive DNA sampling: promise and pitfalls. Ecological applications. 2000;10(1):283-94. 50. Peakall R, Smouse P. GenAlEx 6.5: genetic analysis in Excel. Population genetic software for teaching and research--an update. Bioinformatics (Oxford, England). 2012;28(19):2537-9.

51. Kalinowski ST. hp‐rare 1.0: a computer program for performing rarefaction on measures of allelic richness. Molecular Ecology Notes. 2005;5(1):187-9. 52. Lewontin R, Kojima K-i. The evolutionary dynamics of complex polymorphisms. Evolution. 1960:458-72. 53. Excoffier L, Lischer HE. Arlequin suite ver 3.5: a new series of programs to perform population genetics analyses under Linux and Windows. Molecular ecology resources. 2010;10(3):564-7. 54. Rice WR. Analyzing tables of statistical tests. Evolution. 1989;43(1):223-5. 55. Weir BS, Cockerham CC. Estimating F-statistics for the analysis of population structure. evolution. 1984:1358-70. 56. Jost L. GST and its relatives do not measure differentiation. Molecular Ecology. 2008;17(18):4015-26. 57. Crawford NG. SMOGD: software for the measurement of genetic diversity. Molecular Ecology Resources. 2010;10(3):556-7. 58. Pritchard JK, Stephens M, Donnelly P. Inference of population structure using multilocus genotype data. Genetics. 2000;155(2):945-59. 59. Durand E, Jay F, Gaggiotti OE, François O. Spatial inference of admixture proportions and secondary contact zones. Molecular Biology and Evolution. 2009;26(9):1963-73. 60. Evanno G, Regnaut S, Goudet J. Detecting the number of clusters of individuals using the software STRUCTURE: a simulation study. Molecular ecology. 2005;14(8):2611-20. 61. Earl DA. STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation genetics resources. 2012;4(2):359-61. 62. Jakobsson M, Rosenberg NA. CLUMPP: a cluster matching and permutation program for dealing with label switching and multimodality in analysis of population structure. Bioinformatics. 2007;23(14):1801-6. 63. François O, Durand E. Spatially explicit Bayesian clustering models in population genetics. Molecular Ecology Resources. 2010;10(5):773-84.

64. Spiegelhalter DJ, Best NG, Carlin BP, Van Der Linde A. Bayesian measures of model complexity and fit. Journal of the Royal Statistical Society: Series B (Statistical Methodology). 2002;64(4):583-639. 65. Mantel N. The detection of disease clustering and a generalized regression approach. Cancer research. 1967;27(2 Part 1):209-20. 66. Peakall R, Ruibal M, Lindenmayer DB. Spatial autocorrelation analysis offers new insights into gene flow in the Australian bush rat, Rattus fuscipes. Evolution. 2003;57(5):1182-95. 67. Piry S, Alapetite A, Cornuet J-M, Paetkau D, Baudouin L, Estoup A. GENECLASS2: a software for genetic assignment and first-generation migrant detection. Journal of heredity. 2004;95(6):536-9.

68. Paetkau D, Slade R, Burden M, Estoup A. Genetic assignment methods for the direct, real‐time estimation of

migration rate: a simulation‐based exploration of accuracy and power. Molecular ecology. 2004;13(1):55-65. 69. Rannala B, Mountain JL. Detecting immigration by using multilocus genotypes. Proceedings of the National Academy of Sciences. 1997;94(17):9197-201. 70. Wilson GA, Rannala B. Bayesian inference of recent migration rates using multilocus genotypes. Genetics. 2003;163(3):1177-91. 71. Nunney L, Elam DR. Estimating the effective population size of conserved populations. Conservation Biology. 1994;8(1):175-84. 72. Garza J, Williamson E. Detection of reduction in population size using data from microsatellite loci. Molecular ecology. 2001;10(2):305-18. 73. Cornuet JM, Luikart G. Description and power analysis of two tests for detecting recent population bottlenecks from allele frequency data. Genetics. 1996;144(4):2001-14. 74. Piry S, Luikart G, Cornuet J. BOTTLENECK: a computer program for detecting recent reductions in effective population size using allele frequency data. Journal of Heredity. 1999;90:502-3. PubMed Central PMCID: PMC10.1093/jhered/90.4.502.

75. Dinerstein E, McCracken GF. Endangered Greater One‐horned Rhinoceros Carry High Levels of Genetic Variation. Conservation Biology. 1990;4(4):417-22.

76. Dutta T, Sharma S, Maldonado JE, Wood TC, Panwar H, Seidensticker J. Fine‐scale population genetic structure

in a wide‐ranging carnivore, the leopard (Panthera pardus fusca) in central India. Diversity and Distributions. 2013a;19(7):760-71. 77. Thapa K, Nepal S, Thapa G, Bhatta SR, Wikramanayake E. Past, present and future conservation of the greater one-horned rhinoceros Rhinoceros unicornis in Nepal. Oryx. 2013;47(03):345-51. 78. Blanford WT. The Fauna of British India: Including Ceylon and Burma: Taylor & Francis; 1898. 79. Thapa GB, Weber KE. Actors and factors of deforestation in ‘Tropical Asia’. Environmental conservation. 1990;17(01):19-27. 80. JLD S, C M, SC A, A J, K C. Metapopulation structure of tigers in Nepal. In:Seidensticker J, Christie S, Jackson P In: Seidensticker J, S C, P J, editors. Riding the tiger: tiger conservation in human-dominated landscapes Cambridge: Cambridge University Press; 1999. 81. Nagendra H, Karmacharya M, Karna B. Evaluating forest management in Nepal: views across space and time. Ecology and Society. 2005;10(1). 82. Ceballos G, Ehrlich PR. Mammal Population Losses and the Extinction Crisis. Science. 2002;296(5569):904-7. doi: 10.1126/science.1069349. 83. Haag T, Santos A, Sana D, Morato R, Cullen Jr L, Crawshaw Jr P, et al. The effect of habitat fragmentation on the genetic structure of a top predator: loss of diversity and high differentiation among remnant populations of Atlantic Forest jaguars (Panthera onca). Molecular Ecology. 2010;19(22):4906-21. 84. Sugimoto T, Aramilev VV, Kerley LL, Nagata J, Miquelle DG, McCullough DR. Noninvasive genetic analyses for estimating population size and genetic diversity of the remaining Far Eastern leopard (Panthera pardus orientalis) population. Conservation Genetics. 2014;15(3):521-32. 85. Mondol S, Bruford MW, Ramakrishnan U. Demographic loss, genetic structure and the conservation implications for Indian tigers. Proceedings of the Royal Society of London B: Biological Sciences. 2013;280(1762):20130496. 86. Dutta T, Sharma S, Maldonado JE, Wood TC, Panwar HS, Seidensticker J. Gene flow and demographic history of leopards (Panthera pardus) in the central Indian highlands. Evolutionary applications. 2013;6(6):949-59. 87. Luo S-J, Kim J-H, Johnson WE, Van Der Walt J, Martenson J, Yuhki N, et al. Phylogeography and genetic ancestry of tigers (Panthera tigris). PLoS Biol. 2004;2(12):e442. 88. Reddy PA, Kumaraguru A, Bhagavatula J, Gour DS, Bhavanishankar M, Sarkar MS, et al. Tiger presence in a hitherto unsurveyed jungle of India–the Sathyamangalam forests. Conservation Genetics. 2012;13(3):779-87. 89. Borthakur U, Saini RP, Gupta SS, Jakher R, Das C, Das AK, et al. Noninvasive genetic assessment of population status of tigers (Panthera tigris tigris) in Buxa Tiger Reserve, West Bengal, India. International Journal of Biodiversity and Conservation. 2013;5(1):27-32. 90. Kenney J, Allendorf FW, McDougal C, Smith JL. How much gene flow is needed to avoid inbreeding depression in wild tiger populations? Proceedings of the Royal Society of London B: Biological Sciences. 2014;281(1789):20133337. 91. Chanchani P, Lamichhane B, Malla S, Maurya K, Bista A, Warrier R, et al. Tigers of the Transboundary Terai Arc Landscape: Status, distribution and movement in the Terai of India and Nepal. National Tiger Conservation Authority, Government of India, and Department of National Park and Wildlife Conservation, Government of Nepal NTNC/DNPWC. 2014:3.

92. Safner T, Miller MP, McRae BH, Fortin M-J, Manel S. Comparison of Bayesian clustering and edge detection methods for inferring boundaries in landscape genetics. International Journal of Molecular Sciences. 2011;12(2):865-89.

93. Blair C, Weigel DE, Balazik M, Keeley AT, Walker FM, Landguth E, et al. A simulation‐based evaluation of methods for inferring linear barriers to gene flow. Molecular Ecology Resources. 2012;12(5):822-33. 94. Patil N, Kumar N, Gopalaswamy AM, Karanth KU. Dispersing tiger makes a point. CAMBRIDGE UNIV PRESS 32 AVENUE OF THE AMERICAS, NEW YORK, NY 10013-2473 USA; 2011.

95. Peery MZ, Kirby R, Reid BN, Stoelting R, DOUCET‐BËER E, Robinson S, et al. Reliability of genetic bottleneck tests for detecting recent population declines. Molecular Ecology. 2012;21(14):3403-18. 96. Williamson-Natesan EG. Comparison of methods for detecting bottlenecks from microsatellite loci. Conservation Genetics. 2005;6(4):551-62. 97. Fabbri E, Caniglia R, Galov A, Arbanasid H, Lapini L, Boškovid I, et al. Genetic structure and expansion of golden jackals (Canis aureus) in the north-western distribution range (Croatia and eastern Italian Alps). Conservation genetics. 2014;15(1):187-99. 98. Nagendra H, Pareeth S, Sharma B, Schweik CM, Adhikari KR. Forest fragmentation and regrowth in an institutional mosaic of community, government and private ownership in Nepal. Landscape Ecology. 2008;23(1):41-54. 99. Acharya KP. Twenty-four years of community forestry in Nepal. International Forestry Review. 2002;4(2):149-56. 100. Shrestha TK, Aryal A, Rai RK, Lamsal RP, Koirala S, Jnawali D, et al. Balancing wildlife and human needs: the protected forest approach in Nepal. Natural Areas Journal. 2014;34(3):376-80. 101. GTI. Global tiger recovery program: implemetnation plan 2013-2014. Washington, DC: Global Tiger Initiative Secretariate.; 2013.

Supporting Information

S1 Table. Primers information for tiger specific species and sex identification. bp: base pair; F:

forward; R: reverse

Primer

Name

Sequence (5‘-3‘) Melting

Temp

Product size Reference

TIF ATAAAAAATCAGGAATGGTG 550C 162 bp Bhagavatula& Singh

(2006) TIR TGGCGGGGATGTAGTTATCA 65

0C

AMEL-F

CGAGGTAATTTTTCTGTTTACT

550C 194 bp& 214

bp (male); 214

bp (female)

Pilgrim et al. 2005

AMEL-R GAAACTGAGTCAGAGAGGC 57.30C

CYTB-

SCT-F

AAACTGCAGCCCCTCAGAA

TGATATTTGTCCTCA

72.90C 150 bp Janecka et al. 2008

S2 Table. Thermo-cycling conditions for tiger species and sex identification PCR. PCR:

polymerase chain reaction; min: minute; sec: second; ―x‖ indicates times; F: forward; R: reverse

Steps Species PCR

(TI-F/R)

Sex PCR

(AMEL –F/R)

Cytochrome B PCR

(CYTB-SCT-F/R)

Temperature Time Cycling Temperature Time Cycling Temperature Time Cycling

Initial

denaturation

95°C 15 mins 1X 95°C 15 mins 1X 95°C 15

mins

1X

Denaturation 94°C 30 sec 35X 94°C 30 sec 45X 94°C 15 sec 45X

Annealing 59°C 90 sec 53°C 1 min 55°C 30 sec

Extension 72°C 90 sec 72°C 1 min 72°C 1 min

Final

Extension

72°C 10 mins 1X 72°C 10 mins 1X 72°C 5 mins 1X

Hold 4°C 30 mins 1X 4°C hold 1X 4°C hold 1X

(PCR: polymerase chain reaction; min: minutes; sec: second ―x‖ indicates times; F: forward; R: reverse.)

S3 Table. Genetic variability of 17 candidate microsatellite loci screened and “*” indicates loci

used in this study. NA, number of alleles; Ho, observed heterozygosity.

Locus Allele Size Range (bp) Repeat type Ho NA Source

Microsatellite Primer

FCA205* 102-116 DI 0.78 5 Menotti-Raymond et al.1999

FCA232* 100-108 DI 0.44 3 Menotti-Raymond et al.1999

FCA272 115-127 Di 0.667 6 Menotti-Raymond et al.1999

FCA304* 122-140 Di 0.622 5 Menotti-Raymond et al.1999

FCA391* 122-154 Tetra 0.64 6 Mondol et al.2009a

FCA441* 78-118 Tetra 0.57 4 Mondol et al.2009a

FCA453 69-89 Tetra 0.43 5 Mondol et al.2009a

FCA742 160-184 Tetra 0.55 6 Menotti-Raymond et al.1999

F41 100-172 Tetra 0.429 4 Mondol et al.2009a

F42 152-180 Tetra 0.43 5 Mondol et al.2009a

F53* (165 - 181) Tetra 0.778 5 Menotti-Raymond et al.1999

F85* 155-167 Tetra 0.6 4 Menotti-Raymond et al.1999

PttD5* 200-224 Tetra 0.61 4 Mondol et al.2009a

PttA2 188 - 198 DI 0.51 4 Mondol et al.2009a

Pun82 105-123 DI 0.483 5 Janecka et al. 2008

FCA043* 111-117 DI 0.52 6 Menotti-Raymond et al.1999

FCA008 128-140 DI 0.55 4 Menotti-Raymond et al.1999

S4 Table. Summary results from analysis of molecular variance (AMOVA) for tigers detected

across three populations across the Terai Arc Landscape implemented in program ARLEQUIN

3.5 [53]. df = degree of freedom, P value (α=0.05)

Source of Variation df Sum of Squares Variance Component Percentage of Variation P Value

Among Populations 2 32.155 0.28998 13.72 <0.00

Within Populations 153 279.031 1.82373 86.28 <0.00

Total 155 311.186 2.11371

S5 Table. Summary of pair-wise migration rate (immigration and emigration) between three

populations estimated in Program BayesAss+[70].Net migration rates (immigration-emigration)

were estimated as 0.02 for CNP, -0.08 for SWR, and +0.10 for BNP; ―+‖ indicate migrant receiving

from other population; ―-‖ indicate contributing migrant to other population.

Population

Migrating from

CNP BNP SWR

Migrating

into

CNP 0.98 0.01 0.01

BNP 0.04 0.83 0.13

SWR 0.02 0.02 0.96

S6 Table. Results from program Bottleneck showing the expected and actual numbers of loci

with heterozygosity excess under the respective mutation models, and significance of

heterozygosity excess. IAM: Infinite Allele Model; TPM: Two-Phase Mutation Model; SMM:

Stepwise-Mutation Model. Assuming any mutation model, a Wilcoxon test results with P<0.05

signifies significant heterozygous excess, suggesting that a bottleneck event occurred in CNP only.

Population Heterozygote excess IAM TPM SMM

CNP

Expected 4.53 4.71 4.77

Actual 6 5 3

p value 0.02 0.62 0.80

BNP

Expected 4.59 4.72 4.74

Actual 5 3 3

p value 0.23 0.87 0.87

SWR

Expected 4.33 4.67 4.65

Actual 7 6 6

p value 0.09 0.23 0.23

S1 Fig. Century-wide land-use change detected in the Terai Arc using Anthrome 2.0

datasets[36].

S2 Fig. Magnitude of ∆k± SD (rate of change in the log probability of k; SD: standard deviation)

and Ln P(k) ± SD (posterior probability of the data; SD: standard deviation) as a function of

k(number of sub-populations) detected three and four genetic clusters in the sampled

population following [60].

S3 Fig. STRUCTURE bar plot at k = 2 (top) and k = 4 (bottom) visualizing individual-based

genetic differentiation in tigers across the Terai landscape, Nepal.

S4 Fig. Optimal number of genetic clusters (kmax) based on DIC (Deviance Information Criteria)

for admixture models (CAR and BYM). Both models selected three genetic clusters across the

landscape. Error bars represent standard deviations

S5 Fig. Spatial autocorrelogram for Panthera tigris tigris in the Terai Arc Landscape, Nepal. The

spatial correlogram for tigers (n = 78) shows the genetic correlation coefficient (r) as a function of

geographic distance across defined spatial distance classes. Dashed red lines represent upper (U) and

lower (L) bounds of the null hypothesis based on 9,999 random permutations. Error bars represent

95% confidence intervals about r based on 999 bootstraps.

CHAPTER V

Species, sex and geo-location identification of seized tiger (Panthera tigris tigris) parts in Nepal-

a molecular forensic approach

Dibesh Karmacharya1,7

*, Adarsh M Sherchan1, Santosh Dulal

1, Prajwol Manandhar

1, Sulochana

Manandhar1, Jyoti Joshi

1, Susmita Bhattarai

1, Tarka R Bhatta

1, Nagendra Awasthi

1, Ajay N Sharma

1,

Manisha Bista1, Nawa R Silwal

2, Pravin Pokharel

2, Rom R Lamichhane

3, Netra Sharma

4, Bronwyn

Llewellyn4, Marcella J. Kelly

5, Digpal Gour

6, Lisette Waits

6, Jean M Hero

7, Jane Hughes

7

1 Center for Molecular Dynamics Nepal, Thapathali-11, Kathmandu, Nepal

2 Central Investigation Bureau (CIB), Pillar 4, Nepal Police, Kathmandu, Nepal

3 Bio-Diversity Section, Ministry of Forest and Soil Conservation, Kathmandu, Nepal

4 Environment Team, U.S. Agency for International Development (USAID), Kathmandu, Nepal

5Department of Fish and Wildlife Conservation, Virginia Tech, Blacksburg, Virginia, USA

6 Laboratory for Ecological, Evolutionary and Conservation Genetics, University of Idaho, USA

7School of Environment, Griffith University, Australia

(*Corresponding Author, Email: [email protected])

[This article has been submitted to PLOS One with a revised version based on review obtained: Jan

2018]

Abstract

Tiger (Panthera tigris) populations are in danger across their entire range due to habitat loss,

poaching and the demand of tiger parts. The Bengal tiger (Panthera tigris tigris) is an endangered

apex predator with a population size estimated to be less than 200 in Nepal. In spite of strict wildlife

protection laws, illegal trade in tiger parts has been increasing; and Nepal has become one of the

major sources and transit routes for poached wildlife parts. Identification of wildlife parts is often

challenging for law enforcement officials due to inadequate training and lack of available tools. Here

we describe a molecular forensic approach to gain insight into illegally trafficked tiger parts. We

created Nepal‘s first comprehensive reference genetic database of wild tigers through Nepal Tiger

Genome Project (2011-2013). This database has nuclear DNA microsatellite genotype and sex

profiles, including geo-spatial information, of over 60% (n=120) of the wild tigers of Nepal. We

analyzed 15 putative cases of confiscated poached tiger parts and all were confirmed to be tigers. Ten

were identified as males and five were females, and we determined probable geo-source location for 9

of 14 samples with 6-8 loci of nuclear DNA microsatellite data by combining inferences from three

different assignment methods. Six samples were assigned to Bardia National Park and one of these

was an exact match to a female tiger in our fecal DNA reference database. Two samples were

assigned to Shuklaphanta Wildlife Reserve. One sample was assigned to Chitwan National Park. We

are unable to definitively assign five samples which could be offspring of migrants or samples from

populations outside of Nepal. Our study revealed the western region, particularly Bardia National

Park, is a poaching hotspot for illegal tiger trade across the Terai Arc Landscape in Nepal. We present

scientific evidence to incriminate criminals in a court of law and advocate for the increased use of

molecular forensics in wildlife conservation efforts.

Keywords: Tiger, Wildlife crime, Terai arc landscape, Molecular Forensics, Tiger database

Introduction

The Bengal tiger (Panthera tigris tigris) is the most prevalent and endangered tiger subspecies

residing primarily in Indian subcontinent [1, 2]. In just over a century, the wild tiger population has

dramatically fallen (>97%) and its current population is estimated to be lower than 3,200 [2].

Widespread deforestation, habitat fragmentation and loss, prey depletion, and poaching are the major

causes of tiger population decline [3-6]. Moreover, it has been documented that diseases such as

canine distemper virus (CDV) could threaten fragmented tiger populations [7]. Among these,

poaching and illegal wildlife trade likely present the greatest threat for the survival of the species [8-

10]. With increasing demand for wildlife parts and illicit trade of wildlife, Asia has become one of the

major hubs for wildlife crimes [4, 11, 12]. Nepal is one of the major sources of illegal wildlife parts

[13, 14]. Each year lucrative, illicit wildlife commodities including tiger parts such as skin, bone,

claws, teeth, blood, genitals, and meat are illegally trafficked to Chinese markets [14, 15] and used in

traditional Chinese medicine [15-17]. Wildlife crime in Nepal is often driven by a multitude of factors

including poverty, lucrative financial rewards from illicit wildlife products in black markets, porous

borders, lack of conservation awareness, and poor law enforcement for the prevention of wildlife

crimes [18, 19].

Often the only evidence recovered from the wildlife crime scene is traces of blood, pieces of meat,

skin, fur, bones and other forms of adulterated biological materials. These are virtually impossible to

identify based on morphological analysis alone. Lack of precise identification of confiscated

specimens and species involved often means no convictions in a court of law. Hence, there is an

urgent need to develop advanced forensic techniques to identify seized wildlife specimens to

determine their species and region of origin.

Applications of DNA profiling in wildlife forensics remain rare, although they have been used

successfully in investigations for whales and dolphins conservation [20, 21]; in Canada with mule

deer, white-tailed deer, moose, caribou and black bear [22]; in Italy with poaching cases of wild boar

and wolves [23, 24]; in Africa with large seizures of elephant ivories [25-27] and in India with

seizures of leopard parts [28]. Use of this method in tiger conservation has been rare, though one

study demonstrated feasibility in a tiger bone smuggling case in China [29] and in the illegal sale of

tiger meat from a circus tiger [30]. In Nepal, utility of genetic tools in both population studies and

forensics have not yet been explored for tigers [31], although recently tiger population genetics study

was completed through Nepal Tiger Genome Project (NTGP) (2011-2013) [32].

Genetic tools combined with geo-location baseline data on species can discern geographic origin of

confiscated wildlife parts and their derivatives [22, 25-28, 33, 34]. This kind of information can be

useful for conservation managers to prioritize their efforts in effective anti-poaching activities. In

reviewing success of investments in tiger conservation, best practices in anti-poaching projects have

been those incorporating scientifically sound wildlife monitoring; projects that have been

unsuccessful failed to effectively disseminate data to management and other conservation institutions

[35]. The Government of Nepal (GoN) has asserted that volume of seized tiger parts is high. Nepal is

both a source and a conduit for wildlife parts trafficking into China/Tibet and other areas. Hence there

is a need to utilize new approaches to prevent poaching and illegal trade in tiger parts. The GoN has

proposed establishing a system to better track records of poaching incidences, confiscation and

enforcement of CITES by increasing capacity for identifying animal parts and their derivatives using

effective and modern DNA based forensic tools. [31]

Molecular forensics can provide DNA evidence from illegal wildlife seizures on species, sex and

individuals of poached animals with relatively high accuracy [25, 26, 28, 34, 36-41]. Using a geo-

spatial reference genetic database, we can determine individual or sub-population level source of the

seized parts. Various statistical and computational analyses can assign the most probable geographical

location from which the seized animal parts originated. In Nepal, the main habitat of the Bengal tiger

is the Terai Arc Landscape (TAL) which includes five protected areas (Banke National Park, Bardia

National Park (BNP), Chitwan National Park (CNP), Parsa Wildlife Reserve (PWR) and Suklaphanta

Wildlife Reserve(SWR)) [42, 43].

The primary goal of this study is to use a molecular forensic approach for species, sex and geo-source

location identification of seized tiger parts using our previously developed comprehensive geo-spatial

genetic (DNA) reference database of wild tigers of Nepal [32]. This database now has geo-spatial

information on sex and individual identification for 120 wild tigers of Nepal, which represent 60% of

the estimated 200 tigers in the country. We received 15 confiscated putative tiger parts from the

Central Investigation Bureau (CIB) of Nepal Police and carried out molecular forensic analysis to

determine their species, sex, and probable geo-location source.

Materials and Methods

Source of forensic samples

The laboratory of the Center for Molecular Dynamics Nepal (CMDN) received a total of fifteen

seized putative tiger parts which included thirteen skin pieces and two blood smeared knives (S1 Fig)

from the CIB (S1 Table ). These samples were labeled, photographed and stored with desiccant at

room temperature. Samples were seized from different parts of Nepal from 2014 to 2016.

Geo-spatial baseline genetic database of wild tiger in Nepal

NTGP created Nepal‘s first comprehensive tiger genetic reference database by collecting and

analyzing fecal samples (n=770; CNP=420, BNP=116, SWR=79, PWR and Other Corridors=155)

from the TAL (Fig. 1 and S2 Table), which included all the known habitat of tigers in Nepal ([32];

Thapa et al., in review as in [44]). Collected fecal samples were analyzed for species, sex and

individual identification.

Figure 1 Scat sample collection (n=770; CNP=420, BNP=116, SWR=79, PWR and Other

Corridors=155) sites for tiger baseline genetic database under Nepal Tiger Genome Project

(NTGP, 2011-2013). The estimated tiger population of Nepal was 198 (CNP=120, BNP=50,

SWR=17, PWR= 7, BaNP=4) [45].

Genetic analyses of forensic samples-

DNA extraction

DNA was extracted using DNeasy Blood and Tissue Kit as per manufacturer‘s instruction (Qiagen

Inc., Germany) [46]. A 1 cm2 piece of skin (forensic) sample was cut using a sterile scissors and was

digested overnight at 56 C in 180 μL ATL and 20 μL Proteinase K solution. Extracted DNA was

precipitated with ethanol and purification conducted in a spin column following kit protocol. For

blood smeared knife, samples were taken by swabbing the smeared surface with a sterile cotton swab

saturated with phosphate buffer saline. DNA was extracted using protocol similar to that used for skin

samples. Each batch of DNA extraction included a negative extraction control. Precaution was taken

to avoid contamination. Extracted DNA was stored at -20 °C.

Tiger Identification

Tigers were identified using a PCR assay that used tiger specific mtDNA Cytochrome-b (CYT-B)

primers [47]. A total of 7 μL PCR reaction was prepared containing 3.5 μL of 2X Qiagen multiplex

mastermix, 0.7 μL Q-solution, 200 nM each CYT-B primer sets and 1.5 μL of extracted DNA. The

thermocycling condition was 95 ºC for 15 min followed by 35 cycles of 94 ºC for 30 sec, 59 ºC for 90

sec and 72 ºC for 90 sec with the final extension at 72 ºC for 10 min. Amplified 162 bp target PCR

product was visualized under 1.5% agarose gel electrophoresis (S2 Fig).

Tiger sex identification

Sex of identified tiger samples was determined by amplifying Amelogenin (AMEL) gene of sex

chromosomes developed by Pilgrim et al. [48]. A 7 μL PCR reaction was prepared containing 3.5 μL

of 2X Qiagen multiplex mastermix, 0.7 μL Q-solution, 200 nM each AMEL forward and reverse

primers and 1.5 μL of extracted DNA. The thermocycling condition was 95 ºC for 15 min followed by

45 total cycles of 94 ºC for 30 sec, 53 ºC for 60 sec and 72 ºC for 60 sec with final extension at 72 ºC

for 10 min. Amplified PCR products were visualized in 3% agarose gel electrophoresis. Female

samples yielded a single (214 bp) PCR band, while two bands (194 bp and 214 bp) identified male

samples . Each sample was run in triplicate. Samples that had amplification of X and Y alleles on at

least two out of three replicates were identified as male. Samples that had only X allele amplification

in all three replicates were assigned as female following protocol described by Janecka et al. (2008)

[49].

Individual identification

Individual tigers were identified using a panel of microsatellite markers (n=10) developed from the

domestic cat (Felis catus) and tiger (Panthera tigris tigris) genomes [50-52]. PCR amplification was

carried out in a 7 µL reaction volume containing 3.5 µL of Qiagen multiplex master-mix (Qiagen Inc.,

Germany), 0.7 µL of Q solution, each ten microsatellite primers (FCA391: 0.2 µM, PttD5: 0.07 µM,

FCA232: 0.14 µM, FCA304: 0.07 µM, F85: 0.30 µM, FCA441: 0.14 µM, FCA043: 0.09 µM, and

F53: 0.49 µM) and 2.5 µL of DNA template. The PCR thermocycling condition was 95 ºC for 15 min

with a 10 cycles of touchdown step (94 ºC for 30 sec, initial annealing at 62 ºC reduced by 0.5 ºC in

each cycle for 90 sec and extension at 72 ºC for 60 sec), followed by 25 cycles of 94 ºC for 30 sec, 57

ºC for 90 sec and 72 ºC for 60 sec, and a final extension at 72 ºC for 10 min. Amplified product (0.7

µL) was genotyped by adding 0.3 µL of LIZ-500 size standard in ABI 310 genetic analyzer (Applied

Biosystems, USA). Microsatellite alleles were scored in GENEMAPPER-4.1 (Applied Biosystems,

USA). To finalize the consensus genotypes, a multi-tube approach was used where at least three

identical homozygote PCR results were required for a genotype homozygote call; each allele was

observed in two independent PCRs to record a heterozygous genotype [53]. We calculated probability

of identification (PID) to be 1.5E-06 and probability of identity among siblings (PID(sibs)) to be 3.2E-03

(<0.01) for 8 loci, and achieved an overall genotyping success of 40%. Of all the evaluated

microsatellite markers, eight were selected for genetic analysis of all samples with fairly high overall

PCR amplification success (84%) and genotyping accuracy (82%) (S3 Table). Two additional

microsatellite markers (FCA205 and PttA2) were included in the PCR multiplex but were not

considered in the individual identification as one (Locus FCA205) did not amplify in the majority of

samples and the other (PttA2) was not polymorphic. Similarly, the mean allelic dropout rate was

2.46% and false alleles were around 15.85% (Thapa et al., in review).

Population genetic analysis

The 8 loci microsatellite genotypes from 120 identified individual tigers were used as the reference

baseline data obtained through NTGP [32]. A Bayesian clustering approach was used to determine the

genetic structure in tiger population using STRUCTURE version 2.3.4 software [54, 55]. The

Bayesian clustering was implemented using 2,000,000 Markov chain Monte Carlo (MCMC)

replications after initial 500,000 burn-in with repetitions of the analysis for 10 independent times per

each assumed population (K=1 to 6). This analysis was run using the admixture model with correlated

allele frequencies and was tested under supervised (with LOCPRIOR) and unsupervised (without

LOCPRIOR) learning algorithms. In supervised method, sampling location information consisting of

three geographical locations (N=3) were provided as a priori information, whereas it was not

provided in unsupervised method. The rate of change of likelihood (delta K) value was estimated by

Evanno method [56] to determine the optimal numbers of genetic clusters or populations present in

our reference genotype data for both supervised and unsupervised analysis using STRUCTURE

HARVESTER, a web-version program as described [57]. The percentage of inferred ancestry (Q-

scores) for each samples from 10 independent runs were aggregated into an average Q-scores using

CLUMPP version 1.1.2 software [58].

Forensic assignment of unknown samples

Forensic assignment was conducted using Bayesian clustering analyses with baseline data and

frequency and Bayesian assignment tests using reference populations. A STRUCTURE analysis with

same parameters as described above was performed by incorporating the genotype data from forensic

samples of unknown population of origin (PopInfo=0) and known population samples (PopInfo=1).

The membership of unknown samples to a sub-population was determined based on assignment to

distinct population cluster using percentage of Q-score distribution of each unknown samples. We

interpreted results as an assignment to one of the source populations when average q value scores ≥

70% and classified individuals with < 70% as unassigned and admixed [59]. Geneclass2 [60] was

used to estimate the probability of assignment to each source population with assignment threshold

score of 0.05 using the Paetkau et al 1995 frequency based method [61] with a missing allele

frequency of 0.01 and Rannala and Mountain‘s Bayesian method [62]. We assigned an individual to a

population of origin when the probability of assignment was >0.99 and classified as unassigned when

less than this value. To establish conservative criteria for forensic assignment, we required a sample to

meet the assignment criteria defined above for 2 of the 3 methods. These three methods were chosen

because they have been shown to perform well in accurately assigning individuals in multiple studies

[34, 63, 64].

Results

Genetic analyses

All forensic samples (n=15) were from tigers (S2 Fig). Sex identification revealed males (n=10) and

females (n=5) (S3 Fig) through AMEL gene PCR. There was 100% genotyping success on all forensic

samples across all eight loci (S4 Table), except for one of the blood-smeared knife samples (F-NP-

0012) in which more than half of the loci failed to amplify. An admixed genotype (microsatellite loci)

pattern was observed in F-NP-0012, where three different alleles were detected at three loci FCA391,

FCA232 and FCA043 hence, this sample was excluded from further analyses. By using matching tool

in GenAlEx [65], the genotypes of one of the forensic samples (F-NP-0004, determined as female

(Table 1)) was found to match completely across all loci (n=8) with a female tiger that is in the NTGP

baseline tiger reference database from BNP.

Reference population genetic structure

The supervised approach (with LOCPRIOR) of Bayesian clustering analysis performed in

STRUCTURE predicted two as the most probable number of genetic clusters (K=2) based on Evanno

methodology, where one group consisted of samples from CNP, while the other constituted of

samples from BNP and SWR (Fig 2 and S4 Fig). However, K=3 was the most supported structure

from mean likelihood (mean LnP(K) = -2097.47 and SD=9.6), and q value plots where the three

clusters corresponded to those samples from CNP, BNP and SWR. The latter structure pattern with

three genetic clusters (K=3) was also demonstrated to be the clearest to interpret from Q-score plots

across CNP, BNP and SWR, making an effective reference structure that would be essential for

forensic assignment of unknown samples.

Fig 2 Structure bar plot of (K = 2 to 5) with LOC PRIOR information (supervised algorithm).

The unsupervised approach in STRUCTURE also predicted two as the most probable number of

genetic clusters (K=2) based on Evanno method segregating first cluster consisting of CNP and

another cluster from BNP and SWR (Fig 3 and S5 Fig). However, this model seemed to support K=4

from mean likelihood (mean LnP(K) = -1999.7 and sd=0.34). The CNP population showed

substructures at K=4, but no significant spatial patterns were found. Therefore, we used the supervised

model (with loc prior) with K=3 subpopulation clusters for the population assignment of unknown

forensic samples.

Fig 3 Structure bar plot of (K=2 to 5) without LOC PRIOR information (unsupervised

algorithm)

Forensic assignment of unknown samples

Using K=3 structure of LOCPRIOR model as the reference population structure, the probability

scores of unknown samples were used to assign them at geographical population scale. Based on the

threshold of 70% on Q-scores, eight of the samples were assigned to a specific reference population in

Nepal, while remaining six were determined to be of admixed genetic makeup (Table 1). Among the

eight assigned samples, one sample (F-NP-0009) was assigned to CNP, five samples (F-NP-0004, -

0006, -0010, -0011 and -0013) were assigned to BNP, and two samples (F-NP-0002 and -0014) were

assigned to SWR. For the admixed individuals, 3 showed greater than 50% ancestry with SWR and

most of the remaining ancestry with BNP while the other 3 showed greater than 50% ancestry with

BNP and most of remaining ancestry with SWR (2) or CNP (1). Results from Geneclass analyses

were very similar to STRUCTURE results. Using the frequency-based and Bayesian assignment

methods, 6 samples assigned to BNP, 2 assigned to SWR and 1 assigned to CNP with >99%

probability (Table 1). As observed in the STRUCTURE analyses, 5 samples did not meet our

assignment probability threshold (>99%) and showed admixed ancestry among 2 or more populations

(Table 1). Based on our criteria of requiring 2 or more methods to indicate definitive assignment of

forensic samples (see methods), our final assignments were 6 BNP, 2 SWR and 1 CNP and 5

unassigned samples (Table 1).

Table 1 Species, sex and individual profile of all forensic samples. Species and sex profile of 15

forensic samples. Q-score values from Structure, frequency assignment likelihood and Bayesian

assignment likelihood values from Geneclass2 analyses of 14 forensic samples.

Sample ID Genetic profile Structure q Values

GeneClass2 Assignment

Likelihoods

Geneclass 2 Bayesian

Assignment Likelihoods Forensic

Assigment

Probable

origin Species Sex BNP SWR CNP BNP SWR CNP BNP SWR CNP

F-NP-0001 Tiger Male 0.287 0.629 0.084 0.376 0.611 0.013 0.252 0.747 0.001

Unassigned -

Admixed

BNP/SWR

F-NP-0002 Tiger Male 0.265 0.721 0.015 0 0.999 0 0 0.999 0 SWR Western

Terai

F-NP-0003 Tiger Male 0.461 0.500 0.039 0.669 0.22 0.11 0.807 0.073 0.119

Unassigned -

Admixed

BNP/SWR/CNP

F-NP-0004 Tiger Female 0.817 0.159 0.023 0.998 0.002 0 0.999 0 0 BNP Western

Terai

F-NP-0005 Tiger Female 0.535 0.450 0.016 0.707 0.293 0 0.709 0.29 0

Unassigned -

Admixed

BNP/SWR

F-NP-0006 Tiger Female 0.713 0.261 0.026 0.991 0.008 0 0.997 0.003 0 BNP Western

Terai

F-NP-0007 Tiger Male 0.394 0.577 0.029 0.626 0.35 0.024 0.831 0.137 0.003

Unassigned -

Admixed

BNP/SWR

F-NP-0008 Tiger Female 0.540 0.127 0.334 0.865 0 0.135 0.641 0 0.359

Unassigned -

Admixed

BNP/CNP

F-NP-0009 Tiger Male 0.089 0.178 0.733 0 0 0.999 0 0 0.999 CNP Eastern

Terai

F-NP-0010 Tiger Male 0.815 0.166 0.020 0.991 0.009 0 0.984 0.016 0 BNP Western

Terai

F-NP-0011 Tiger Male 0.916 0.066 0.018 0.999 0 0 0.999 0 0 BNP Western

Terai

F-NP-0013 Tiger Male 0.884 0.095 0.021 0.999 0 0 0.999 0 0 BNP Western

Terai

F-NP-0014 Tiger Male 0.225 0.756 0.019 0 0.999 0 0.001 0.999 0 SWR Western

Terai

F-NP-0015 Tiger Female 0.669 0.233 0.098 0.999 0 0 0.999 0 0 BNP Western

Terai

F-NP-0012 Tiger Male - - - - - - - - - Genotyping failed -

*Bold highlighting indicates that a sample meets defined criteria for assignment ≥70% for Structure

and ≥99% for Geneclass 2

Discussion

Wildlife crime is a serious trans-national threat to biodiversity in South Asia/Indian subcontinent.

There is an immediate need to develop tools that can assist in tracking and ultimately preventing

wildlife crime which is driven by organized criminal syndicates that control the burgeoning and

highly lucrative illicit trade in wildlife parts [66]. CIB has been actively involved in the investigation

of wildlife crime in all regions of Nepal and has had major successes in bringing the perpetrators to

justice by closely working with other national and international stakeholders. Higher seizure cases of

tiger body parts from Mid/Far western Nepal compared to other regions also indicate that this might

be one of the major trafficking routes for illegal wildlife trade in Nepal. The recently established

molecular forensic capability in Nepal has provided them with a new tool in the fight against wildlife

crime in Nepal.

The use of forensic tools plays an important role in strengthening law enforcement efforts to address

wildlife crime and biodiversity conservation. It has already been successfully used in some countries

to improve investigation of wildlife cases, identification of poaching hotspots and trafficking routes,

as well as prosecution of wildlife criminals by connecting individuals to crime scenes [24-26, 28, 36,

37, 66].

TAL is a biologically diverse habitat. Once alluvial grasslands and subtropical deciduous forests of

TAL has supported the highest recorded density of tigers in the world [67]. Poaching now remains a

great threat to tiger survival in this landscape. Dwindling tiger numbers in some protected areas call

into question the current protection strategy. Tiger populations in the western part of the TAL may

already have reached a dangerously low tipping-point, and further threat may push the sub-population

towards local extinction, as has occurred for rhino population across Babai River floodplain in BNP

[18, 68].

Application of molecular (DNA) forensics is relatively new and the rapidly evolving technology is

gaining acceptance in wildlife conservation efforts [69, 70]. The information obtained through the

forensic genetics is more robust and comprehensive than conventional methods [36, 71]. To fulfill the

urgency to address location of poached tigers, we created a wild tiger geo-spatial, genetic baseline

database through the NTGP [32]. This has provided valuable information on wildlife crime ―hot

spots‖ and presented scientific evidence to incriminate criminals in a court of law. We identified one

tiger skin (F-NP-0004; female) that had an exact DNA genotyping profile (PID (unrelated)1.5x10-6

; PID

(sibs) 3.2x10-3

) with one of our female reference NTGP tigers from BNP. We collected scats of that

individual in the winter of 2012; this particular individual was probably poached as recent as one to

two years ago. Of all the forensic evidence we analyzed, most were traced to BNP (Table 1),

highlighting this area as a wildlife crime hotspot. This is crucial information for Nepal‘s law

enforcement officials. After dissemination of these results (Table 1), the concerned stakeholders,

including the Nepal Police and the Nepal Army increased their patrols and intelligence activities

particularly in the BNP region. As a result, a few of the criminal networks involved in wildlife crime

in that area were discovered and necessary actions were taken supporting long term wildlife crime

prevention [72, 73]. Our analysis confirmed that 15 of 15 suspected tiger parts were from tigers and

10 of 15 samples were male.

The molecular assay we have used in our baseline tiger genetic database and forensic samples have

high probability of identification (PID; 1.5E-06) and probability of identification among siblings (PID

(sibs); 3.2E-03) with eight used highly polymorphic microsatellite loci [32]. Similarly, our species and

sex identification PCR assay was highly effective (Thapa et al., in review). Other studies have shown

that assignment tests can be very effective and accurate when pairwise Fst values between source

populations exceed 0.05 as is the case in this study [63] and (Thapa et al., in review). Using our

established criteria, we were able to definitely assign 9 of 14 individuals to a source population in

Nepal. Five individuals were unassigned and showed admixed genetic profiles between CNP, SWR

and BNP. These individuals may have been offspring of migrants. Previous research detected 3

migrants from SWR to BNP and showed higher gene flow from SWR into BNP than BNP to SWR.

This is also reflected in the STRUCTURE q value plots that show a larger proportion of admixed

individuals in BNP (Figure 2). Thus, these admixed individuals are more likely to have been poached

in BNP than SWR (Thapa et al., in review). Another possibility is that the samples came from across

the border in India where we did not have reference samples. To test this hypothesis, we generated

Nei‘s pairwise genetic distance values for all individuals and built a UPGMA phylogenetic tree to

search for outlier samples (S1 Supporting Information). Two of the unassigned samples (F-NP-0001

and -0008) in this analysis did not cluster with the 3 main groups and may represent individuals from

unsampled source populations in India.

Our tiger genetic database covers about 60% of estimated overall tiger population of Nepal and we do

not have any data from tiger habitats in India that share forest connectivity with Nepal. Hence, the

tiger source geo-location data and sub-population assignment analysis could have been more accurate

if we had a more comprehensive database of sampling including samples from the Indian tiger

population of TAL. We acknowledge that some of the samples assigned to regions in Nepal could

have been poached just across the border in India where tiger allele frequencies are likely similar.

There is now an urgent demand to have trans-boundary collaboration between tiger habitat countries

[74, 75] to build and to share standardized tiger genetic data so that comprehensive and cross-

referencing anti-poaching genetic database can be developed. We also recommend a molecular

forensic data validation process to evaluate accuracy of the technique and integrate this in courts of

law as admissible evidence in wildlife crime cases where this is not currently in place by conducting

multi-year scat surveys and building a robust baseline genetic database of wild tigers not only in

Nepal but throughout the tiger range countries in South Asia. There is an ongoing effort to have court

of law in Nepal accept DNA based evidence. To certain extent some DNA based evidence is accepted

(species identification). We are working with all the relevant stakeholders, including United States

Fish and Wildlife Services (USFWS) and INTERPOL to bring awareness on the utility of Molecular

Forensics in Nepal. This work is a major part of our effort in developing molecular forensic tool in the

fight against wildlife crime in the region.

Conclusions

The world is dealing with an unprecedented spike in illegal wildlife trade, threatening to overturn

decades of conservation gains. Wildlife trade is the fourth most lucrative illegal trade in the world

after drugs, human trafficking and the arms trade [76]. To combat wildlife crime through providing

scientifically accurate and reliable information, Nepal now has developed baseline geo-location based

genetic database of wild tigers and molecular forensic capacity. This database can be effectively used

to identify any tiger up to sub-population and/or individual level. With efforts like NTGP, developing

countries like Nepal have been making the significant progress in building capacity to gather

important biodiversity information on various wildlife species. This forensic study instructs the

crucial initiatives and activities to effective enforcement of laws for wildlife protection and

conservation by contributing a powerful tool in the fight against poaching and wildlife crime.

Acknowledgments

We would like to thank the Department of National Parks and Wildlife Conservation and the Ministry

of Forest and Soil Conservation (Nepal) for giving us permission to conduct the first genetic study of

tigers in Nepal. We would also like to thank CIB and the Nepal Police team for providing us with the

opportunity to analyze the samples. All our collaborators had important contributions in providing

valuable feedback while designing our study and analyzing data.

References

1. WWF. Overview and Threats of Tiger. Text. http://www.worldwildlife.org/species/tiger. World Wildlife Fund (WWF). 2016. 2. Perry R. The world of the tiger. Cassell and Comp. Ltd. London.1964. 264 p. 3. Seidensticker J. Saving wild tigers: a case study in biodiversity loss and challenges to be met for recovery beyond 2010. Integrative zoology. 2010;5(4):285-99. 4. Dinerstein E, Loucks C, Wikramanayake E, Ginsberg J, Sanderson E, Seidensticker J, et al. The fate of wild tigers. . BioScience. 2007;57(6):508-14.

5. Barber‐Meyer SM, Jnawali SR, Karki JB, Khanal P, Lohani S, Long B, et al. Influence of prey depletion and human

disturbance on tiger occupancy in Nepal. Journal of Zoology. 2013;289(1):10-8.

6. Kenney JS, Smith JL, Starfield AM, McDougal CW. The long‐term effects of tiger poaching on population viability. . Conservation Biology 1995 Oct 1;9(5):1127-33. 1995;9(5):1127-33. 7. Gilbert M, Soutyrina SV, Seryodkin IV, Sulikhan N, Uphyrkina OV, Goncharuk M, et al. Canine distemper virus as a threat to wild tigers in Russia and across their range. Integrative zoology. 2015;10(4):329-43. 8. McMurray CA. Testimony of Claudia A. McMurray to the Committee on Natural Resources of U.S. House of Representatives (5 March). Hearing on: poaching American security: impacts of illegal wildlife trade, Washington, D.C., USA. 2008. 9. Goodrich J, Lynam A, Miquelle D, Wibisono H, Kawanishi K, Pattanavibool A, et al. 2015. Panthera tigris. The IUCN Red List of Threatened Species 2015: e.T15955A50659951. http://dx.doi.org/10.2305/IUCN.UK.2015-2.RLTS.T15955A50659951.en. . 2015. 10. Brown P, Davies B. Trading to extinction. Dewi Lewis Publishing, UK. 2014. 11. Wyler LS, Sheikh PA. International illegal trade in wildlife: Threats and U.S. Policy. Congressional Research Service, USA. 2013. 12. Stoner SS, Pervushina N. Reduced to Skin and Bones Revisited: An Updated Analysis of Tiger Seizures from 12 Tiger Range Countries (2000–2012). TRAFFIC, Kuala Lumpur, Malaysia. 2013. 13. Yi-Ming L, Zenxiang G, Xinhai L, Sung W, Niemela J. Illegal wildlife trade in the Himalayan region of China.: Kluwer Academic Publishers, Netherlands; 2000. 901-18 p. 14. Moyle BJ. The black market in China for tiger products. Global Crime 2009;10(1-2):124-43. 15. Ellis R. Tiger bone & rhino horn: the destruction of wildlife for traditional Chinese medicine. Island Press. 2013. 16. Mainka SA, Mills JA. Wildlife and traditional Chinese medicine: supply and demand for wildlife species. Journal of zoo and wildlife medicine. 1995:193-200. 17. Gratwicke B, Mills J, Dutton A, Gabriel G, Long B, Seidensticker J, et al. Attitudes toward consumption and conservation of tigers in China. PloS one. 2008;3(7):e2544. 18. Martin EB, Martin C. Insurgency and poverty: recipe for rhino poaching in Nepal. Pachyderm. 2006;41:61-73. 19. Roe De. Conservation, crime and communities: case studies of efforts to engage local communities in tackling illegal wildlife trade. IIED, London. 2015. 20. Palumbi A, Cipriano F. Species identification using genetic tools: the value of nuclear and mitochondrial gene sequences in whale conservation. Journal of Heredity. 1998;89(5):459-64. 21. Baker C, Cipriano F, Palumbi S. Molecular genetic identification of whale and dolphin products from commercial markets in Korea and Japan. Molecular Ecology. 1996;5(5):671-85. 22. Ogden R, Dawnay N, McEwing R. Wildlife DNA forensics—bridging the gap between conservation genetics and law enforcement. Endangered Species Research. 2009;9(3):175-95. 23. Lorenzini R. DNA forensics and the poaching of wildlife in Italy: a case study. Forensic science international. 2005;153(2-3):218-21. 24. Caniglia R, Fabbri E, Greco C, Galaverni M, Randi E. Forensic DNA against wildlife poaching: identification of a serial wolf killing in Italy. Forensic science international Genetics. 2010;4(5):334-8. 25. Wasser SK, Shedlock AM, Comstock K, Ostrander EA, Mutayoba B, Stephens M. Assigning African elephant DNA to geographic region of origin: applications to the ivory trade. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(41):14847-52. 26. Wasser SK, Mailand C, Booth R, Mutayoba B, Kisamo E, Clark B, et al. Using DNA to track the origin of the largest ivory seizure since the 1989 trade ban. Proceedings of the National Academy of Sciences. 2007;104(10):4228-33. 27. Wasser S, Brown L, Mailand C, Mondol S, Clark W, Laurie C, et al. Genetic assignment of large seizures of elephant ivory reveals Africa’s major poaching hotspots. Science. 2015;349(6243):84-7. 28. Mondol S, Sridhar V, Yadav P, Gubbi S, Ramakrishnan U. Tracing the geographic origin of traded leopard body parts in the indian subcontinent with DNA-based assignment tests. Conservation biology : the journal of the Society for Conservation Biology. 2015;29(2):556-64. 29. Xu YC, Li B, Li WS, Bai SY, Jin Y, Li XP, et al. Individualization of tiger by using microsatellites. Forensic science international. 2005;151(1):45-51. 30. Wan QH, Fang SG. Application of species-specific polymerase chain reaction in the forensic identification of tiger species. Forensic science international. 2003;131(1):75-8.

31. Bajimaya S, Karki JB, Pant G, Kafle MN, Jnawali SR, Bajracharya S, et al. Tiger Conservation Action Plan for Nepal 2008-2012. Department of National Parks and Wildlife Conservation (DNPWC), Ministry of Forests and Soil Conservation, Government of Nepal. Available: http://www.wwfnepal.org/?191044/Tiger-Conservation-Action-Plan2008-2012. 2007. 32. NTGP/CMDN. Nepal Tiger Genome Project - Final Report 2014, Center for Molecular Dynamics-Nepal. Center for Molecular Dynamics-Nepal, 2014. 33. PalsbØLl PJ, Berube M, Skaug HJ, Raymakers C. DNA registers of legally obtained wildlife and derived products as means to identify illegal takes. Conservation Biology. 2006;20(4):1284-93. 34. Manel S, Berthier P, Luikart G. Detecting wildlife poaching: identifying the origin of individuals with Bayesian assignment tests and multilocus genotypes. Conservation biology. 2002;16(3):650-9. 35. Gratwicke B, Seidensticker J, Shrestha M, Vermilye K, Birnbaum M. Evaluating the performance of a decade of Save The Tiger Fund's investments to save the world's last wild tigers. Environmental Conservation. 2007;34(03):255-65. 36. Alacs EA, Georges A, FitzSimmons NN, Robertson J. DNA detective: a review of molecular approaches to wildlife forensics. Forensic science, medicine, and pathology. 2010;6(3):180-94. 37. Budowle B, Van-Daal A. Extracting evidence from forensic DNA analyses: Future molecular biology directions. Biotechniques. 2009;46(5):339. 38. Nagata J, Aramilev VV, Belozor A, Sugimoto T, McCullough DR. Fecal genetic analysis using PCR-RFLP of cytochrome b to identify sympatric carnivores, the tiger Panthera tigris and the leopard Panthera pardus, in far eastern Russia. Conservation Genetics. 2005;6(5):863-6. 39. Karmacharya DB, Thapa K, Shrestha R, Dhakal M, Janecka JE. Noninvasive genetic population survey of snow leopards (Panthera uncia) in Kangchenjunga conservation area, Shey Phoksundo National Park and surrounding buffer zones of Nepal. BMC research notes. 2011;4(1):516. 40. Gurgul A, Radko A, Słota E. Characteristics of X-and Y-chromosome specific regions of the amelogenin gene and a PCR-based method for sex identification in red deer (Cervus elaphus). Molecular biology reports. 2010;37(6):2915-8. 41. Palomares F, Godoy J, Píriz A, O'Brien S. Faecal genetic analysis to determine the presence and distribution of elusive carnivores: design and feasibility for the Iberian lynx. Molecular ecology. 2002;11(10):2171-82. 42. Wikramanayake E, McKnight M, Dinerstein E, Joshi A, Gurung B, Smith D. Designing a conservation landscape for

tigers in human‐dominated environments. . Conservation Biology. 2004;18(3):839-44.

43. Kanagaraj R, Wiegand T, Kramer‐Schadt S, Anwar M, Goyal SP. Assessing habitat suitability for tiger in the fragmented Terai Arc Landscape of India and Nepal. . Ecography 2011 Dec 1;34(6):970-81. 2011;34(6):970-81. 44. Thapa K, Manandhar S, Bista M, Shakya J, Sah G, Dhakal M, et al. Assessment of genetic diversity, population structure, and gene flow of tigers (Panthera tigris tigris) across Nepal's Terai Arc Landscape. [Journal Article under review at PlosOne]. In press 2017. 45. Dhakal M, Karki M, Jnawali SR, Subedi N, Pradhan NMB, Malla S, et al. Status of Tigers and Prey in Nepal: Department of National Parks and Wildlife Conservation, Kathmandu, Nepal; 2014. 46. Qiagen. DNeasy Blood & Tissue Handbook 2006. Available from: http://diagnostics1.com/MANUAL/General_Qiagen.pdf. 47. Bhagavatula J, Singh L. Genotyping faecal samples of Bengal tiger Panthera tigris tigris for population estimation: a pilot study. BMC genetics. 2006;7:48. 48. Pilgrim K, McKelvey K, Riddle A, Schwartz M. Felid sex identification based on noninvasive genetic samples. Molecular Ecology Notes. 2005;5(1):60-1.

49. Janečka J, Jackson R, Yuquang Z, Diqiang L, Munkhtsog B, Buckley‐Beason V, et al. Population monitoring of snow leopards using noninvasive collection of scat samples: a pilot study. Animal Conservation. 2008;11(5):401-11. 50. Menotti-Raymond M, David VA, Lyons LA, Schäffer AA, Tomlin JF, Hutton MK, et al. A genetic linkage map of microsatellites in the domestic cat (Felis catus). Genomics. 1999;57(1):9-23.

51. Sharma R, Stuckas H, Moll K, Khan I, Bhaskar R, Goyal S, et al. Fourteen new di‐and tetranucleotide

microsatellite loci for the critically endangered Indian tiger (Panthera tigris tigris). Molecular ecology resources. 2008;8(6):1480-2. 52. Menotti-Raymond MA, David VA, Wachter LL, Butler JM, O’Brien SJ. An STR forensic typing system for genetic individualization of domestic cat (Felis catus) samples. J Forensic Sci. 2005;50(5):1061-70. 53. Gill P, Ivanov PL, Kimpton C, Piercy R, Benson N, Tully G, et al. Identification of the remains of the Romanov family by DNA analysis. Nature genetics. 1994;6(2):130-5. 54. Pritchard JK, Stephens M, Donnelly P. Inference of population structure using multilocus genotype data. Genetics. 2000;155(2):945-59. 55. Hubisz MJ, Falush D, Stephens M, Pritchard JK. Inferring weak population structure with the assistance of sample group information. Molecular ecology resources. 2009;9(5):1322-32. 56. Evanno G, Regnaut S, Goudet J. Detecting the number of clusters of individuals using the software STRUCTURE: a simulation study. Molecular ecology. 2005;14(8):2611-20. 57. Earl DA. STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation genetics resources. 2012;4(2):359-61. 58. Jakobsson M, Rosenberg NA. CLUMPP: a cluster matching and permutation program for dealing with label switching and multimodality in analysis of population structure. Bioinformatics. 2007;23(14):1801-6.

59. Wultsch C, Waits LP, Kelly MJ. A Comparative Analysis of Genetic Diversity and Structure in Jaguars (Panthera onca), Pumas (Puma concolor), and Ocelots (Leopardus pardalis) in Fragmented Landscapes of a Critical Mesoamerican Linkage Zone. PloS one. 2016;11(3):e0151043. 60. Piry S, Alapetite A, Cornuet J-M, Paetkau D, Baudouin L, Estoup A. GENECLASS2: a software for genetic assignment and first-generation migrant detection. Journal of heredity. 2004;95(6):536-9. 61. Paetkau D, Calvert W, Stirling I, Strobeck C. Microsatellite analysis of population structure in Canadian polar bears. Molecular ecology. 1995;4(3):347-54. 62. Rannala B, Mountain JL. Detecting immigration by using multilocus genotypes. Proceedings of the National Academy of Sciences. 1997;94(17):9197-201. 63. Latch EK, Dharmarajan G, Glaubitz JC, Rhodes OE. Relative performance of Bayesian clustering software for inferring population substructure and individual assignment at low levels of population differentiation. Conservation Genetics. 2006;7(2):295-302. 64. Hauser L, Seamons T, Dauer M, Naish K, Quinn T. An empirical verification of population assignment methods by marking and parentage data: hatchery and wild steelhead (Oncorhynchus mykiss) in Forks Creek, Washington, USA. Molecular Ecology. 2006;15(11):3157-73. 65. Peakall R, Smouse PE. GENALEX 6: genetic analysis in Excel. Population genetic software for teaching and research. Molecular ecology notes. 2006;6(1):288-95. 66. Iyengar A. Forensic DNA analysis for animal protection and biodiversity conservation: A review. Journal for Nature Conservation. 2014;22(3):195-205. 67. Sunquist M. Tigers: Ecology and behavior. Pp. 19-33 In: Tigers of the World: The science, politics and conservation of Panthera tigris. (eds. R. Tilson & P.J. Nyhus), Elsevier, San Diego, USA. . 2010. 68. Thapa K. Situation Analysis of Babai Valley in Bardia National Park- Preliminary Report. WWF Nepal. 2006. 69. Hsieh HM, Chiang HL, Tsai LC, Lai SY, Huang NE, Linacre A, et al. Cytochrome b gene for species identification of the conservation animals. Forensic science international. 2001;122(1):7-18. 70. Schwartz MK, Luikart G, Waples RS. Genetic monitoring as a promising tool for conservation and management. Trends in ecology & evolution. 2007;22(1):25-33. 71. Ogden R. Unlocking the potential of genomic technologies for wildlife forensics. Molecular Ecology Resources. 2011;11(s1):109-16. 72. Pariyar K. Banjara involvement a new trend in tiger poaching: Police. Republica. 2016 Jan 13, 2016. 73. Shahi P. Poaching still a major threat to tigers. Kathmandu Post. 2016 Feb 12, 2016. 74. Singh SK, Aspi J, Kvist L, Sharma R, Pandey P, Mishra S, et al. Fine-scale population genetic structure of the Bengal tiger (Panthera tigris tigris) in a human-dominated western Terai Arc Landscape, India. PloS one. 2017;12(4):e0174371. 75. Gour DS, Reddy PA. Need of transboundary collaborations for tiger survival in Indian subcontinent. Biodiversity and conservation. 2015;24(11):2869-75. 76. Dalberg W. Fighting illicit wildlife trafficking: A consultation with governments. WWF International, Gland, Switzerland, 2012.

Supporting information

S1 Fig. Seized tiger skin sample (F-NP-0011) provided by Central Investigation Bureau (CIB) of

Nepal

S2 Fig. A 1.5% agarose gel electrophoresis image of tiger species identification PCR targeting

Cytochrome B region of mitochondrial DNA.

S3 Fig. A 3% agarose gel electrophoresis image of sex identification PCR of tiger forensic samples

targeting Amelogenin gene. Females show single band at 214 bp and males show two bands at 194 bp

and 214 bp.

S4 Fig. Delta K plot of Structure HARVESTER result demonstrating the optimal value (K=3) under

LOCPRIOR model.

S5 Fig. Delta K plot of Structure HARVESTER result demonstrating the optimal value (K=4) under

the model without LOCPRIOR information.

S6 Fig. Phylogenetic tree (UPGMA) generated from Nei‘s genetic distance using 8 nuclear DNA

microsatellite loci for 120 reference tiger samples and 14 forensic samples. Samples are colored based

on sampling locations. Green represents CNP, blue represents BNP, and red represents SWR samples.

Forensic samples are in black. Clusters or clades are labeled accordingly.

S1 Table Forensic samples provided by the Central Investigation Bureau (CIB) of Nepal

Sample Regions of seizure Sample type Year of seizure

F-NP-0001 Far Western Nepal Skin 2015

F-NP-0002 Far Western Nepal Skin 2015

F-NP-0003 Far Western Nepal Skin 2015

F-NP-0004 Mid Western Nepal Skin 2015

F-NP-0005 Mid Western Nepal Skin 2015

F-NP-0006 Mid Western Nepal Skin N/A

F-NP-0007 Far Western Nepal Skin N/A

F-NP-0008 Eastern Nepal Skin 2014

F-NP-0009 Mid Western Nepal Skin 2014

F-NP-0010 Far Western Nepal Skin 2016

F-NP-0011 Far Western Nepal Skin 2016

F-NP-0012 Far Western Nepal Blood smeared knife 2016

F-NP-0013 Far Western Nepal Blood smeared knife 2016

F-NP-0014 Mid Western Nepal Skin 2016

F-NP-0015 Mid Western Nepal Skin 2016

N/A= Not available

S2 Table Proportions of scat sampled and tigers identified from those scats across each study sites

from NTGP.

CNP BNP SWR Others Total

Scat collected 420 116 79 155 770

Tigers

identified

69 32 19 1 120

S3 Table Summary of PCR amplification success, genotyping accuracy and genotyping error rates for

8 microsatellite loci for all processed tiger samples (n = 401) collected in three protected areas

(Chitwan National Park, Bardia National Park, and Suklaphanta Wildlife Reserve) to build baseline

tiger genetic database.

Locus All Samples (n=401)

PCR GA ADO FA

FCA391 75.67 88.59 0.16 10.52

PttD5 88.43 88.82 1.12 10.48

FCA232 87.50 77.02 2.06 23.34

FCA304 94.00 87.46 0.00 10.94

FCA043 93.59 80.26 2.72 17.79

F53 65.84 84.62 0.27 13.09

F85 73.86 73.42 7.94 19.01

FCA441 94.85 74.73 5.40 21.69

Mean 84.22 81.87 2.46 15.86

SD 10.97 6.33 2.85 5.25

PCR, % polymerase chain reaction amplification success; GA, % genotyping accuracy; ADO, %

allelic dropout; FA, % false alleles

S4 Table. Genotype data on 8 microsatellite loci of the identified tiger forensic samples

M. Loci

S. ID

FCA391 PttD5 FCA232 FCA304 F85 FCA043 F53 FCA441

F-NP-0001 152 152 199 199 99 101 121 121 161 161 115 115 0 0 111 111

F-NP-0002 152 152 199 211 101 101 121 125 157 157 115 115 168 168 115 123

F-NP-0003 152 152 199 199 101 101 125 125 165 165 115 119 164 172 115 115

F-NP-0004 152 152 207 211 101 101 121 125 165 165 115 115 160 172 115 119

F-NP-0005 152 156 199 199 101 101 121 125 157 165 115 117 160 164 119 119

F-NP-0006 152 152 211 211 101 101 121 125 165 165 115 123 164 172 111 119

F-NP-0007 152 152 199 199 101 101 121 121 165 165 115 115 164 172 115 115

F-NP-0008 152 152 199 207 101 101 121 123 153 153 129 129 172 172 107 107

F-NP-0009 144 148 199 199 99 99 121 121 0 0 115 115 180 180 111 115

F-NP-0010 152 152 207 211 101 101 121 125 165 165 117 123 164 172 119 119

F-NP-0011 152 152 207 211 99 101 125 125 157 157 115 115 160 160 107 119

F-NP-0013 152 152 207 211 101 101 125 125 0 0 115 115 0 0 107 119

F-NP-0014 152 152 199 211 101 101 125 125 161 161 119 119 0 0 115 115

F-NP-0015 152 152 207 211 101 101 121 121 165 165 129 129 160 172 107 115

S5 Table. Values of Mean log-likelihood and Delta K for each assumed K (K=1 to 6) of both

supervised (with LOC PRIOR) and unsupervised (without LOC PRIOR) models .

K

with LOC PRIOR without LOC PRIOR

Mean LnP(K) Delta K Mean LnP(K) Delta K

1 -2467.34 -2467.44

2 -2191.11 44.90 -2191.80 2723.10

3 -2097.47 1.92 -2097.70 1.05

4 -2022.26 1.67 -1999.70 130.05

5 -1993.51 0.68 -1946.75 26.50

6 -1923.44 -1886.49

S1 Supporting Information. Methodology and Results of Distance-based phylogenetic analysis.

S1 Supporting Information.

Methodology and Results of Distance-based phylogenetic analysis

Phylogenetic inference of microsatellite genotype data

Nei‘s standard genetic distance measure (Ds) was utilized to calculate pair-wise genetic distance

between samples using their microsatellite genotype data to infer a phylogenetic relationship among

samples [1, 2]. A 10,000 bootstrap on loci was performed to construct the distance matrix, which was

used to reconstruct topology of phylogenetic tree from UPGMA clustering algorithm [3]. The Ds

distance matrix calculation and UPGMA tree reconstruction was performed in POPULATIONS

1.2.32 (http://bioinformatics.org/~tryphon/populations) package for Linux. The reconstructed

phylogenetic tree was visualized and drawn using FigTree version 1.4.2

(http://beast.bio.ed.ac.uk/figtree) software.

Result of Phylogenetic analysis

The topology of the phylogenetic tree demonstrated evidence of geo-location based population

clusters of the tiger population (S6 Fig). However, some samples collected from one population got

assigned into other population cluster. A total of nine samples were assigned into population other

than their known collected population site. Four samples that were collected in BNP and two in SWR

fell under CNP clades. Three samples from BNP got assigned into SWR clade. The CNP population

prominently had two major and one minor clusters, and BNP were grouped into one major and one

minor clusters. All CNP clusters were closely placed with two of the BNP clusters, while the SWR

cluster formed a very separate group from both CNP and BNP clusters in the tree topology. There

were evidences of admixture samples which were placed in clusters of different populations. All

forensic samples (n=14) were positioned based on their genetic distance relatedness to individual

reference samples in the tree. Based on this, three forensic samples (F-NP-0001 (male), -0008

(female), and -0009 (male)) formed a complete out-group from the tree, which reflected their distant

relatedness with our reference tiger genetic database. Among remaining eleven samples, six (F-NP-

0004, -0005,-0010, -0011, -0013, and -0015) were closely related to BNP samples while five (F-NP-

0002, -0003, -0006, -0007, and -0014) were closely related to SWR samples.

References

1. Nei M. Genetic distance between populations. American naturalist. 1972:283-92.

2. Nei M, Tajima F, Tateno Y. Accuracy of estimated phylogenetic trees from molecular data.

Journal of Molecular Evolution. 1983;19(2):153-70.

3. Sokol RR, Michener CD. A statistical method for evaluating systematic relationships, Univ.

Kansas Science Bulletin. 1958;28:1409-38.

CHAPTER VI

Gut microbiome and their putative metabolic functions in fragmented Bengal

tiger population of Nepal

Dibesh Karmacharya1,7

*, Prajwol Manandhar1, Sulochana Manandhar

1, Santosh Dulal

1, Nagendra P.

Awasthi1, Adarsh M. Sherchan

1, Ajay N. Sharma

1, Jyoti Joshi

1, Manisha Bista

1, Shailendra

Bajracharya1, Netra Sharma

2, Bronwyn Llewellyn

2, Lisette P. Waits

3, Kanchan Thapa

4, Marcella J.

Kelly4, Momchilo Vuyisich

5, Shawn R. Starkenburg

5, Jean-Marc Hero

6, Jane Hughes

7, Claudia

Wultsch8, Laura Bertola

9, 10, Nicholas M Fountain-Jones

11, Amit K. Sinha

1

1Center for Molecular Dynamics Nepal, Thapathali-11, Kathmandu, Nepal

2Environment Team, U.S. Agency for International Development-Nepal

3Department of Fish and Wildlife Sciences, University of Idaho, Idaho, USA

4Department of Fish and Wildlife Conservation, Virginia Tech, Blacksburg, Virginia, USA

5Applied Genomics, Los Alamos National Lab, New Mexico, USA

6School of Science & Education, University of the Sunshine Coast, Queensland, Australia

7School of Environment, Griffith University, Queensland, Australia

8Sackler Institute for Comparative Genomics, American Museum of Natural History, New York, USA

9Department of Biology, City College of New York, New York, USA

10Institute of Environmental Sciences, Leiden University, Leiden, The Netherlands

11Department of Veterinary Population Medicine, University of Minnesota, Minnesota, USA

(*Corresponding Author, Email: [email protected])

[This art icle has been submitted to Nature Micr obiome in Jan 2018]

Abstract

Bengal tigers (Panthera tigris tigris) serve a pivotal role as an apex predator in forest ecosystems. To

increase our knowledge on factors impacting the viability and health of this endangered species, we

studied the gut microbiomes in 32 individual Bengal tigers from three geographically separated

protected areas (Chitwan National Park, Bardia National Park and Suklaphanta Wildlife Reserve) in

Nepal, using noninvasive genetic sampling methods. Gut microbiome influence the immune system,

impact various physiological functions, and modulate metabolic reactions, that ultimately impact the

host health, fitness, behavior, digestion and development. Overall analysis of beta diversity revealed

significant differences in the composition of microbial communities across the populations and

bacterial communities clustered based on their geographic locations with moderate levels of overlap.

Microbial clustering across all areas was significant with pairwise tests showing significant

differences between Chitwan and the other two protected areas, but no differences between Bardia

and Suklaphanta. Predictive metagenome functional content analysis (PICRUSt) revealed a higher

functional enrichment and diversity in the Chitwan population and the lowest enrichment and

diversity in Suklaphanta. The Chitwan population contained a lower proportion of microbiome

associated with immune system and infectious diseases, but higher proportions that are associated

with xenobiotic biodegradation/metabolism, cellular processing and signaling, and neurodegenerative

diseases compared to the other two protected areas. We observe the tiger population structure, gut

microbiome profile and associated functional metabolic categories are correlated, with geographically

most separated Chitwan and Suklaphanta population having the most distinct and different host

genotype and microbiome profiles. These findings underscore the importance of conservation and

management programs for tigers and other wildlife species that seek to maximize genetic diversity at

different hierarchical levels, including hosts as well as their microbial communities.

Introduction

Gut microbiota is a complex community of microorganisms in the intestinal tract that has co-evolved

with the host 1,2

playing an important role in maintaining the host‘s health. Gut microbial communities

shape the immune system, impact various physiological functions, and modulate metabolic reactions,

that ultimately impact the host health, fitness, behavior, digestion and development 3-5

. The

composition of gut microbiota is largely determined by numerous intrinsic and extrinsic factors such

as the host‘s environment, health status, genotype, dietary habits, age, sex, social relationships and

disease prevalence 6-10

. The composition of gut microbiota may also change responding to

environmental stresses such as the loss and fragmentation of habitat 11-14

. Surrounding anthropogenic

pressure could be an important modulator of the gut microbiomes and could influence abundance of

various metabolites which are associated with metabolism of cellulose, phenolics, organic acids,

simple sugar and lipids 15

.

In the last decade, spurred by technological advances in DNA sequencing, multiple studies have

described gut microbiota of various terrestrial and aquatic species, including humans, primates,

whales and other mammals 16-18

. Disturbances in the bacterial microbiota might result in

immunological dysregulation that may underlie disorders such as inflammatory bowel disease,

Crohn's disease, and ulcerative colitis 19,20

. The mammalian immune system which appears to control

microbes, in fact, might be controlled by the microbes themselves 21

. By stimulating the immune

system and the development of gut structure, gut microbes play a crucial role in the regulation of host

health by aiding in the defense against invading pathogens and providing nutritional benefit to the

host such as production of short chain fatty acids and vitamin B12 22

. The overall balance in the

composition of the gut microbial community, as well as the presence or absence of key microbiome

species capable of affecting specific responses, is important in ensuring homeostasis or lack thereof at

the intestinal mucosa and beyond. The mechanisms through which microbiota exert beneficial or

detrimental influences remain largely undefined, but include elaboration of signaling molecules and

recognition of bacterial epitopes by both intestinal epithelial and mucosal immune cells 23

. As

microbial communities inhabiting wildlife species affect host health, nutrition, physiology and

immune systems, gut microbiome research has several significant implications for wildlife

conservation and management 7,24

. Therefore, it is recommended to include wildlife microbiome

studies into comprehensive wildlife conservation and management programs, especially for species of

conservation concern 25

.

Globally, the population of wild tigers is declining dramatically due to widespread habitat loss and

fragmentation, prey depletion, illegal hunting and various infectious diseases 26-29

. Within Nepal,

habitat loss and fragmentation have forced extant tigers to divide into distinct geographically

separated populations in Nepal, which has been extensively studied using long-term field data 30

and

noninvasive genetic sampling [Thapa et. al., 2017 in review]. The Bengal tiger‘s main habitat in

Nepal is restricted to five protected areas along the Terai Arc Landscape (TAL), including Chitwan

National Park (CNP), Parsa Wildlife Reserve (PWR), Bardia National Park (BNP), Banke National

Park (BaNP), and Suklaphanta Wildlife Reserve (SWR) 31,32

. The TAL has experienced significant

land use changes in the recent past 27,32

. Human settlements surround and encroach into tiger habitat

degrading natural areas and potentially increasing levels of environmental stress for tigers. This

region has also experienced severe socio-political unrest, which included 10 year civil war during the

Maoist insurgency, that has negatively impacted fragile ecosystems with weakened wildlife

conservation programs 33,34

. Considering the degree of environmental degradation in conjunction with

habitat loss and fragmentation over the last century, it is vital to take a multidimensional approach to

managing the health of wild tiger subpopulations.

The extent to which habitat loss and fragmentation alter the gut microbiome and in turn impact the

health of endangered wildlife is largely unknown. Small isolated wildlife populations may not only

have low genetic diversity but also have lower gut microbial diversity with altered functionality that

could adversely impact the health of these animals and potentially increase the risk of local extinction.

Red colubus monkey, an endangered species, residing around human settlements seemed to have

reduced gut microbial diversity compared to a population found in a wild habitat 12

. In our previous

work we identified 120 individual tigers using eight microsatellite markers and found that tigers in

SWR had the lowest genetic diversity and were the most isolated in terms of gene flow compared to

the other parks. Tigers from CNP and BNP had similar levels of genetic diversity even though CNP is

geographically distant from BNP and SWR [Thapa et.al., under review]. Based on this, we

hypothesized that SWR might have the least diverse microbiome compared to the other tiger

populations. However, as humans can substantially perturb the microbiota of wildlife35

we expect the

tigers in the CNP may have a unique microbial community as this park receives much higher human

visitation compared to the other parks. The aim of our study is to examine structural and functional

diversity of gut microbial communities in tiger populations of TAL (Fig 1).

As part of the Nepal Tiger Genome Project (NTGP) 36

, we conducted one of the largest scale

microbiome surveys of a wild carnivore spanning three populations with different degrees of

connectivity and human visitation. We found that even though microbiome alpha diversity was

similar across tiger populations, there were significant differences in microbial phylogenetic and

taxonomic composition and beta diversity, and these differences were potentially important in terms

of metabolic and immune function. This increased our knowledge of the gut microbiome and can

contribute towards the development of a more comprehensive strategy to conserve and manage wild

tiger populations occurring across fragmented landscapes.

Fig 1. Scat sample collection sites for tiger baseline genetic database under Nepal Tiger

Genome Project (NTGP, 2011-2013). We identified 120 individual tigers using 8 microsatellite

markers from TAL (SWR=19, BNP=32, CNP=69). A total of 70 tiger scat samples from 32 identified

individual tigers (CNP=12; BNP=12; SWR=8) were randomly selected for gut microbiome analysis.

Results

Composition of core tiger gut microbiota across sites

The metagenomic sequencing of tiger scat and soil samples (tiger, n = 70; soil, n = 8) targeting the

hypervariable V4 region of 16S rRNA gene generated a total of 4,385,688 sequences, among which

tiger samples consisted of 2,985,814 sequences and soil samples consisted of 1,399,874 sequences.

For 70 tiger samples from 32 individuals, the mean number of sequences per sample was 42,654

(range: 1,614 – 95,553). Similarly for 8 soil samples, the mean number of sequences per sample was

174,984 (range: 134,283 – 235,580). The bacterial community based characterized Operational

Taxonomic Unit (OTU), with ≥ 97% nucleotide sequence identity, differed significantly between soil

and tiger samples (Beta diversity PERMANOVA; Unweighted Unifrac: pseudo F=12.69, P=0.001;

Weighted Unifrac: pseudo F=15.52, P=0.001) (Fig. 2). Tigers‘ core microbiota was similar in

composition across the three regions (Fig. 3). Overall, the most dominant gut microbiota of the tiger

samples were the Phyla Proteobacteria (~37%), Firmicutes (~27.7%), Bacteroidetes (~18.6%),

Fusobacteria (~13.4%), and Actinobacteria (~2.7%) (Table 1). These taxa formed the core gut

microbiome of tiger, which is similar in composition with findings reported for a number of other

mammals 17,37-39

.

Table 1. Core gut microbiome biodiversity in Bengal tigers. The abundance of ten most prevalent

gut microbiome at Operational Taxonomic Unit (OTU). Bacteria phyla detected in Bengal fecal

samples are summarized for three major protected areas (CNP=Chitwan National Park; BNP=Bardia

National Park; SWR=Suklaphanta Wildlife Reserve) within the Terai Arc Landscape of Nepal.

OTU ID CNP BNP SWR Total

Proteobacteria 41.8% 40.6% 28.4% 36.9%

Firmicutes 23.8% 23.1% 36.1% 27.7%

Bacteroidetes 26.5% 14.8% 14.3% 18.5%

Fusobacteria 4.3% 19.2% 16.7% 13.4%

Actinobacteria 3.09% 1.26% 3.71% 2.69%

Tenericutes 0.10% 0.56% 0.09% 0.25%

TM7 0.02% 0.11% 0.23% 0.12%

Unassigned Other 0.09% 0.11% 0.06% 0.09%

Verrucomicrobia 0.11% 0.004% 0.01% 0.04%

All Other Categories 0.03% 0.04% 0.16% 0.08%

Fig 2. Principal Coordinate Analysis (PCoA) of soil (n = 7; CNP = 3, BNP = 4) and tiger fecal

samples (n = 32; CNP = 12, BNP = 12, SWR = 8) - microbiome profiles for soil are distinct from

fecal samples indicating there is unlikely to be cross contamination between the two.

Fig 3. Gut microbiome biodiversity- Relative abundance of genera of tiger gut

microbiome across three national parks (CNP, BNP, SWR) in Nepal.

Alpha-diversity measures of gut microbiome

Overall there was no significant difference in alpha diversity (within population) of the microbiome

across all three populations independent of the index used (Chao1 and ACE metrics 40,41

, Shannon‘s

index 42

, Simpson‘s index 43

, Inverted Simpson and Fisher‘s indexes 44

) (Fig. 4). Analysis of each

index indicated that BNP was marginally less diverse in terms of species richness with CNP having

marginally higher alpha diversity (Fig. 4).

Fig 4. Alpha-diversity in tiger gut microbial communities is similar across different regions

studied.

Beta-diversity of the tiger gut microbiome across sites

Our analysis of beta diversity (between population) revealed significant differences in the

composition of microbial communities across the tiger populations. Canonical analysis of principal

coordinates of weighted UniFrac distances indicated that bacterial communities clustered based on

their geographic locations with moderate levels of overlap (Fig. 5). One-way PERMANOVA on

weighted UniFrac distances found that the clustering across all areas was significant (pseudo F =

3.086, P = 0.0058) with pairwise tests showing significant differences between CNP and the other two

protected areas (CNP vs BNP: pseudo t = 1.944, P = 0.006; CNP vs SWR: pseudo t =1.9942, P =

0.0071), but no differences between SWR and BNP (t = 0.782, P = 0.606) (Fig 5a). Furthermore,

there was greater beta phylogenetic diversity between CNP compared to SWR (PERMDISP, pseudo t

= 2.72, P = 0.015), but not between CNP and BNP (PERMDISP pseudo t = 1.92, P = 0.083) or BNP

and SWR (PERMDISP pseudo t = 0.79, P = 0.797). The CAP model helped to confirm these results

with samples correctly assigned to each respective group most of the time (57%; null allocation

success = 0.33) (Fig. 5b).

PERMANOVA found significant clustering (pseudo F = 2.712, P = 0.0003) and similar patterns when

compared pairwise (CNP vs SWR: pseudo t = 2.056, P = 0.003; CNP vs BNP: t =1.512, P = 0.005;

BNP vs SWR: pseudo t = 1.114, P = 0.216). CNP had higher diversity than other parks (PERMDISP,

CNP vs BNP, pseudo t = 2.36, P = 0.031, CNP vs SWR: pseudo t = 3.511, P = 0.001), however, there

was no difference in beta-diversity between BNP or SWR (PERMDISP, pseudo t = 1.21, P = 0.27).

The CAP model could correctly assign samples to each population most of the time (63%, null

allocation success = 0.33).

Fig 5. Canonical analysis of principal (CAP) coordinates showing differences in microbiome

composition across three national parks. PERMANOVA of both phylogenetic composition

(weighted UniFrac distance) (a) and taxonomic composition (Bray-Curtis distance) (b) similarity

measures illustrated that CNP had differing microbial composition and overall greater beta diversity.

Statistical analysis for abundant microbe taxa

Analysis of differences in abundance, based on the F test (with Benjamini and Hochberg control for

false discovery rate 45

) identified significant differences in Fusobacteria and TM7 (both with adjusted

P values of 0.018, fdr at 0.017) across the three regions. There was separation in the microbiome

profiles and abundances among samples collected at the three geo-locations, with BNP samples

forming a ―bridge‖, statistically, between relatively well differentiated samples collected in SWR and

CNP. Specifically, the phylum Fusobacteria separated samples collected in CNP from a combined

group of samples from BNP and SWR (adjusted P = 0.01, fdr = 0.028). The abundance of the phylum

Thermi was significantly different in SWR samples from samples collected in BNP and CNP

(adjusted P value=0.072, fdr = 0.168).

This inference motivated a more in depth analysis with consensus sequences obtained for twenty of

the most prevalent genera (based on presence/absence). A statistical analysis with F-test and a

Benjamini and Hochberg correction for false discovery identified significant differences in

Comamonas, Collinsella, and Fusobacterium across CNP, BNP, and SWR with significant adjusted P

values and acceptable rates of false discovery (fdr) (Table 2).

Table 2. Statistical analysis for most abundant microbial phyla and genera in gut microbiomes found

in three sub-populations of tigers in Nepal.

Phylum Genus

Abundance

Adj P

False

discovery

rate (fdr) CNP BNP SWR

Proteobacteria Comamonas 2.47% 0.36% 0.26% 0.03 0.024

Actinobacteria Collinsella 0.61% 0.61% 2.93% 0.03 0.024

Fusobacteria Fusobacterium 3.49% 16.73% 15.79% 0.109 0.064

Predictive metabolic functions associated with tiger gut microbiota

The PICRUSt analysis using all obtained OTU suggested that there is divergence in predicted

functional categories among the gut microbiomes across populations of CNP, BNP and SWR (P value

< 0.05) based on 12 functional categories (Table 3). Pairwise Welch‘s t-test performed to test

statistical differences in mean proportion of functional categories in combinations of two populations

revealed 2 categories among CNP Vs BNP (P value < 0.05) (Fig 6b), and 23 categories among CNP

Vs SWR (p value < 0.05) (Fig 6a), while functional categories did not differ significantly between

BNP and SWR.

Table 3 Predictive functional profiling of microbial communities using multi-group comparative

PICRUSt analysis (P-value corrected < 0.05).

Functional categories P-values

P-values

(corrected)

CNP BNP SWR

mean

rel.

freq.

(%)

std.

dev.

(%)

mean

rel.

freq.

(%)

std.

dev.

(%)

mean

rel.

freq.

(%)

std. dev.

(%)

Metabolism of Other

Amino Acids

2.46E-

05 1.01E-03 1.7586 0.1840 1.6343 0.1300 1.5277 0.1463

Lipid Metabolism

5.11E-

05 1.05E-03 3.3852 0.4784 3.0074 0.2283 2.9102 0.2805

Xenobiotics

Biodegradation and

Metabolism

2.37E-

04 3.23E-03 2.6593 0.7133 2.1849 0.4225 1.9963 0.3467

Metabolism of

Terpenoids and

Polyketides

4.61E-

04 4.72E-03 1.8282 0.2626 1.6218 0.1454 1.6080 0.1672

Digestive System

9.37E-

04 7.68E-03 0.0381 0.0179 0.0288 0.0116 0.0211 0.0130

Neurodegenerative

Diseases

1.38E-

03 9.45E-03 0.2326 0.0977 0.1657 0.0805 0.1401 0.0770

Excretory System

3.60E-

03 2.11E-02 0.0288 0.0077 0.0265 0.0049 0.0217 0.0077

Transport and Catabolism

4.38E-

03 2.24E-02 0.3019 0.0999 0.2527 0.0633 0.2222 0.0656

Amino Acid Metabolism

6.04E-

03 2.75E-02 9.8148 0.6008 9.2886 0.6104 9.4261 0.4714

Circulatory System

6.79E-

03 2.78E-02 0.0206 0.0193 0.0094 0.0160 0.0054 0.0136

Replication and Repair

1.13E-

02 4.23E-02 7.5835 0.7039 7.6596 0.7293 8.1723 0.5950

Nucleotide Metabolism

1.19E-

02 4.05E-02 3.5190 0.3695 3.6291 0.3409 3.8212 0.2764

The Welch‘s t-test between CNP and SWR identified higher proportions of predicted functional

categories related to ―amino acid metabolism‖, ―lipid metabolism‖, ―xenobiotics biodegradation and

metabolism‖, and ―metabolism of terpenoides and polyketides‖ in CNP samples. SWR samples had

presence of comparatively higher proportions of ―energy metabolism‖, ―carbohydrate metabolism‖,

―metabolism of cofactors and vitamins‖ and ―nucleotide metabolism‖ categories. The BNP samples

contained lower proportions of ―lipid metabolism‖ and ―metabolism of terpenoids and polyketides‖

than the CNP samples.

Fig 6. Mean proportions of predictive metabolic functional categories between tiger populations

based on pair-wise comparison- functional metabolic categories are more diverse between CNP

and SWR than CNP and BNP. (a) Significant metabolic functional categories identified between

CNP and SWR. (b) Significant metabolic functional categories identified between CNP and BNP.

Overall significant differences were observed in predictive metabolic functions over the tiger

population of the three study regions. The pair-wise relationship between samples based on functional

analysis corresponded to the pair-wise relation between samples based on microbiome structure

(PROCRUSTES, m2 = 0.72, P value = 0.0009). This analysis further underscores the geo-location

specificity with the sample set from BNP overlapping with well-separated sample sets from CNP and

SWR.

Discussion

Fragmented population and variation in gut microbiome in tigers

Low population numbers in some tiger populations and increased levels of habitat loss and

fragmentation may contribute towards lowering of genetic diversity in host species, which in turn can

also adversely impact the microbiomes coevolving within their bodies. In the TAL region of Nepal,

there has been a 97% increase in agriculture and settlement areas of the past 200 years and forested

areas decreased by 47% between the 19th and 20

th centuries (Thapa et al 2017, in review). In our

study, we found some significant differences in gut microbial composition in three geo-spatially

separated tiger population. These differences were highest between CNP and SWR tiger populations

which are geographically most separated (Fig. 1). Differences in habitat including differences in

prey composition and tiger densities, as well as interactions across the fragmented population of these

national parks, might have role in shaping such microbiota composition and abundance. For example,

the microbiome diversity results mirror the results from neutral microsatellite diversity with SWR

having lower diversity than CNP (Thapa et.al., 2017 under review) and support our hypothesis that

SWR would have the lowest levels of microbiome diversity. Our study was also able to find that

there is a notable differences in microbiota diversity observed between CNP and other two protected

areas, as opposed to the differences observed between BNP and SWR, giving an indication that the

microbiota diversity in CNP is unique from that of BNP and SWR. This could be due to the reason

that CNP has more human visitor turnover and/or activity than other two protected areas which we

had hypothesized. Also, the other reason could be that habitat fragmentation in TAL has separated

CNP far away from BNP and SWR, while BNP and SWR still having connections through wildlife

corridors between them. Although this study is first of its kind in wild tigers, a study in an

endangered primate showed a direct correlation between higher habitat fragmentation with reduced

gut microbiota diversity as well as richness with some profound implications on health and long-term

viability of the species 12

.

Gut microbiome and its potential implication on health of tigers

Of the microbiome found in tigers, Proteobacteria was highly abundant in all tiger populations [CNP

(42%), BNP (41%) and SWR (28%)]. Proteobacteria are gram negative bacteria which include a

wide variety of species that are commensal and part of normal gut flora of higher organisms as well as

many true pathogens, such as Escherichia, Shigella, Salmonella, Yersinia, Buchnera, Ha emophilus,

Vibrio, Pseudomonas, and Helicobacter 46

. Helicobacter has been linked to chronic gastritis that can

lead to increased morbidity and mortality among captive cheetah around the globe 47

. Higher

abundance of Proteobacteria has been associated with increased incidences of gastrointestinal diseases

among human and animals like cats and dogs 48-51

.

Firmicutes were also abundant in tiger gut [CNP (24%), BNP (23%) and SWR (36%)]. Some of the

notable pathogens under this Phyla are Bacilli and Clostridia 52

. Firmicutes constitute one of the

predominant phyla (57 % of community) found in the gut of wild mammal species 53

. In Nambian free

ranging cheetahs, it was found to be the dominant phylum constituting 68.5% of total gut microbiota

54. In humans, Firmicutes accounts for more than 80% of bacterial population in colon

8. They were

observed to be dominant in both wild (76%) and captive animals (61%) in sea lions 55

and occupied

the largest portion of the bacterial profile (65–79%) in non-human primates 56

. Firmicutes was also

the predominant phylum (88%) in South American and Sub-antarctic fur seals 57

. Firmicutes along

with Bacteroidetes have predominantly been shown to be involved in energy resorption and

metabolism 58

. Increased ratio of Firmicutes and Bacteroidetes has been linked to potential

development of diabetes and obesity in human and other mammals 59,60

.

Although speculative, in view of studies done on humans, the pathogenic roles and activities of the

three significant virulent bacteria found in tigers (Comamonas, Collinsella, and Fusobacteria) may

have correlation with functional activities in the immune and digestive system and prevalence of

diseases such as neuro-degenerative diseases and cancer 61

.

CNP tiger microbiome had significantly higher (2.47%) presence of Comamonas than BNP (0.36%)

and SWR (0.26%) population (Table 2). Comamonas are aerobic, gram-negative, motile, oxidase-

positive bacilli which usually has low virulence potency and infrequently cause human disease 62

.

Comamonas kerstersii has been occasionally reported as causative agent of intra-abdominal infections

resulting peritonitis secondary to appendix rupture in human cases 62,63

.

SWR population had higher (2.93%) Collinsella than CNP (0.61%) and BNP .(61%). Collinsella

aerofaciens is the most abundant bacterium in the human intestine present in about 1010

cells/ gm of

faeces, and is found in more than 90% of human intestines 64

. A reduced abundance of this species has

been correlated with more severe symptoms in patients suffering from irritable bowel syndrome 65

.

Collinsella is often known to elevate alpha-aminoadipic acid, asparagine and pro-inflammatory

cytokine IL-17A which are known to cause diseases associated with immune system such as

rheumatoid arthritis 66

. Furthermore, Collinsella in combination with Fusobacterium nucleatum

fadA activity, alters gut permeability and disease severity on gastrointestinal and immune diseases 66-

68.

BNP and SWR tiger population had significantly greater presence of Fusobacterium than CNP

population. Fusobacterium is a gram-negative bacterium that ferment carbohydrates or amino acids

thereby producing various organic acids. Fusobacterium necrophorum has been recognized as a true

animal pathogen 69

. It is associated with liver abscesses in cattle, foot rot in many domestic animals,

calf diphtheria and necrotic lesions in the oral cavity 70,71

. It is also closely associated with

development of colon cancer 72

.

Gut microbiome and functional metabolic implication on tiger population

PICRUSt based predictive metabolic functionalities in tiger population revealed a consistent pattern

with higher functional enrichment and diversity in the CNP tiger population for the most categories

and the lowest enrichment and diversity in SWR population. We found significant differences in 2

functional categories among CNP Vs BNP (Fig 6b), and 23 categories among CNP Vs SWR (Fig 6a),

while functional categories did not differ significantly between BNP and SWR (Fig 6). Predictive

metabolic profile is just a rough indicator of possible functional implication of microbiota present. We

observe the tiger population structure, gut microbiome profile and associated functional metabolic

categories are correlated, with geographically most separated CNP and SWR population having the

most distinct and different host genotype and microbiome profiles.

This further highlights the necessity of more comprehensive systems biology based approach to study

conservation status of the species with a focus on monitoring and maintaining genetic diversity and a

healthy gut microbial composition. We are further investigating various other factors that might

influence gut microbiome and its influence on tiger's survival, including role of anthropogenic

proximity and activities. Any conservation effort to preserve highly endangered tiger needs to adopt a

holistic conservation approach with focus on the inter-dependencies between host genetic

differentiations and gut microbial diversity.

Methods

Methods for host genetic analysis

Genetic database of wild tiger in Nepal and fecal DNA sampling

NTGP created Nepal‘s first comprehensive tiger genetic reference database by collecting and

analyzing fecal samples (n = 770) from the TAL (Fig. 1), which included all the known habitat of

tigers in Nepal (Thapa et al in Review; NTGP project report). The TAL has a sub-tropical monsoonal

climate and mixed deciduous vegetation ranging from alluvial floodplain grasslands communities to

Climax Sal (Shorea robusta) forests and includes five protected areas, among which Suklaphanta

National Park (SWR, 28°50′25″N 80°13′44″E), Bardia National Park (BNP, 28°23′N 81°30′E), and

Chitwan National Park (CNP, 27°30'0.00" N 84°40'0.12" E) are the major tiger habitats (Fig. 1).

Putative tiger fecal (scat) samples were collected from protected areas and connecting wildlife

corridors across the TAL-Nepal (Thapa et al in Review). Ninety-eight grid cells each measuring 15 X

15 km (225 km2, sampling unit) were sampled using opportunistic field surveys. A few grams from

the upper surfaces of the scat were removed and stored at room temperature in sterile 2-ml vials filled

with DETs buffer dimethyl sulphoxide saline solution; 73

at 1:4 volume scat-to-solution ratio

following field sampling protocols by Wultsch, et al. 74

.

DNA extraction, species identification, and individual identification

We extracted DNA from scat samples using a commercially available QIAmp DNA Stool Kit

(QIAGEN Inc., Germany) following the manufacturer‘s instructions. Each batch of DNA extraction

included a negative extraction control. Extracted DNA was stored at -20 °C. Tigers were identified

using PCR assay that used tiger specific mtDNA Cytochrome-b (CYT-B) primers 75

. Individual tigers

were identified by microsatellite analysis using a panel of eight microsatellite markers developed from

the domestic cat (Felis catus) and tiger genomes 76-78

as described in Thapa et al. [In Review].

Gut microbiome analysis

We randomly selected a total of 70 scat samples from 32 unique individual tigers (n = 12, CNP; n =

12, BNP; n = 8, SWR) for gut microbiome analysis. Soil samples were also collected from two of the

study sites (n = 3, CNP; n = 5, BNP) with the goal to profile soil microbiomes to assess cross-

contamination between soil and fecal samples occurred.

Bacterial DNA was isolated from tiger fecal and soil samples using PowerSoil DNA Isolation Kit

(MoBio, Qiagen, Carlsbad, CA). DNA quality was checked by gel electrophoresis (mostly >10 kbps

fragments), and DNA concentration was measured by Qubit (Invitrogen, Carlsbad, CA). We

completed microbial community profiling (identification and composition) by amplifying and

sequencing the hyper-variable region (V4) of the 16S rRNA from both tiger scat and soil control

samples using a modified version of the protocol presented in Caporaso et. al 2012 79

, adapted for the

Illumina MiSeq platform. Using a two-step polymerase chain reaction (PCR), we amplified the V4

region of the 16S rRNA using the ‗universal‘ bacterial/archaeal primer pairs (515F and 806R) linked

to the forward and reverse Illumina flow cell adapter sequences. PCR was carried out in two steps,

both using the 2X KAPA HiFi HotStart ReadyMix (KAPA Biosystems/Roche, USA) and cycling at

initial denaturation at 950C/30sec followed by 95

0C/30sec, 55

0C/30sec, 72

0C/30sec. Post cycling,

samples were incubated at 720C for 5 min, followed by a hold at 4

0C. The first PCR was conducted in

25 cycles, adding a 6 bp barcode sequence to enable multiplexing. The second PCR was conducted in

8 cycles to amplify the PCR products and add the remaining full-length Illumina adapters. We

purified the resulting PCR products using Agencourt AMPure XP beads (Beckman Coulter, Brea,

CA), quantified with Qubit (Invitrogen, USA), normalized, and pooled prior to sequencing. Paired-

end sequencing (2 x 300bp) was completed on Illumina MiSeq (Illumina, Inc., San Diego, CA), using

V3 600 cycle kits according to the manufacturer‘s instructions.

After sequencing and de-multiplexing, we filtered all reads by quality (q>29 across 50% of the read

length with no ambiguous N base calls) and length (>75 bp). A custom Perl script was written to

execute several analysis modules of QIIME, version 1.9.1 80

. First, we joined raw paired-end Illumina

fastq files by fastq-join. We discarded all OTU containing less than 10 sequences. We chose the

cluster centroid for each OTU as the OTU representative sequence and taxonomically assigned each

sequence using homologous searches to 16S reference sequences found in the Greengenes database 81

at greater than or equal to 96% sequence identity. To construct a phylogenetic tree of the OTU

representative sequences, we aligned sequences using PyNAST, version 1.2.2 82

against an existing

alignment of the Greengenes database. Post alignment and construction of phylogenetic trees was

completed using FastTree, version 2.1 83

.

Alpha-diversity, beta-diversity estimates, and relative abundance analysis of each taxonomic group

were performed after rarefaction was applied with even sub-sampling of 10,000 sequences per

sample. The abundances of OTUs were normalized based on proportion and OTUs with very low

variability (1e-05

) were filtered out. Microbial diversity within (alpha diversity) and between tiger

subpopulations (beta diversity) were obtained and visualized with QIIME and the phyloseq package in

R 84,85

. We assessed alpha diversity using several metrics (Chao1, ACE, Shannon, Simpson,

InvSimpson, Fisher) 40-44

. Statistical evaluation of differential abundance was done with F test

supplemented in the mt function in phyloseq. The resulting P values were adjusted for multiple

comparisons using Benjamin and Hoechberg‘s False discovery method (Fig. 6).

To test if gut microbiota diversity was significantly differentiated across different study sites (beta

diversity), we employed Permutational Multivariate Analysis of Variance (PERMANOVA) 86

,

canonical analysis of principal coordinates (CAP) 87

, and permutational tests of homogeneity of

dispersions (PERMDISP) 88

. To test for differences in both UniFrac and Bray-Curtis similarity

distances, we performed a one-way PERMANOVA and used pair-wise contrasts to examine

differences between sites. We visualized analyzed compositional differences using CAP and DPCoA

(detrended principal coordinate analysis) and also used the CAP discriminant analysis to validate the

PERMANOVA results (i.e., how distinct was each site in multivariate space) by assessing allocation

success using the ‗leave-one-out‘ procedure 87

. PERMDISP was used to compare beta diversity

between sites for both metrics 88

and to test if differences detected by PERMANOVA were likely due

to differences in-group dispersions. All of the above analyses were conducted in PRIMER- E

PERMANOVA+89

.

Predictive metabolic functions associated with tiger gut microbiota:

We used PICRUSt (Phylogenetic Investigation of Communities by Reconstruction of Unobserved

States) 90

to predict functional roles played by the tiger gut microbiome communities. PICRUSt

predicts metabolic and functional profiles of a microorganism based on known functional roles of its

closely related microorganism 90

. It utilizes existing information from Integrated Microbial Genomes

(IMG) database 91

, which contains annotation of gene content and 16S copy number data of reference

bacterial and archaeal genomes. Then by implementing extended ancestral-state reconstruction

algorithm, the taxonomic composition and phylogenetic information of the observed OTUs are used

in estimating the comprehensive metagenome of the microbiome community classifying their

metabolic and functional categories in the KEGG Orthology (KO) classification scheme 92

. The

PICRUSt predictions were subjected to statistical analyses with Statistical Analysis of Metagenomic

Profiles (STAMP) 93

for identifying and characterizing significant functional categories across three

subpopulations CNP, BNP, and SWR. We conducted multiple group statistical tests with ANOVA

and pair-wise statistical tests using Welch‘s t-test to test for statistical differences in mean proportion

of functional categories among subpopulations. The p-value was adjusted by applying the Benjamini-

Hochberg false discovery rate (FDR) method to correct for multiple hypotheses testing. We conducted

Procrustes 94

analysis using QIIME to test correlations on beta-diversity obtained for gut microbiome

and predictive microbiome functionality contents using Bray-Curtis distance metrics.

Acknowledgments

We would like to thank the Department of National Parks and Wildlife Conservation and the Ministry

of Forest and Soil Conservation (Nepal) for giving us permission to conduct the first genetic study of

tigers in Nepal. We would also like to thank our collaborators at the Los Alamos National lab, Griffith

University and the University of Idaho for providing valuable technical support and resources. And

finally, this work would not have been possible without dedication and hard work from our team at

the Center for Molecular Dynamics Nepal.

Author contributions

Study Concept- DK

Research Design- DK, PM,SM, LPW, JH and AKS were involved in designing the study

Field study design & sample collection- DK, KT, MJK, and MB designed field sample collection plan

and were involved in sample collection.

Laboratory analysis- SM, NPA, AMS, ANS, JJ, MV and SRS performed the laboratory work.

Data analysis- DK, PM, NMFJ, CW, SB and AKS did bioinformatics, statistical and GIS data

analysis.

Manuscript preparation- DK, SD, NS, BL, LPW, MV, JMH and CW worked on manuscript

preparation.

Additional Information

Competing interests

The funder supporting this research had no role in study design, data collection, interpretation and

decision to submit the work for publication.

Data Availability

All data related to the study is included in the manuscript and are available through the npj Biofilms

and Microbiomes website.

REFERENCES

1 Ley, R. E. et al. Evolution of mammals and their gut microbes. Science 320, 1647-1651,

doi:10.1126/science.1155725 (2008).

2 Quercia, S. et al. From lifetime to evolution: timescales of human gut microbiota adaptation.

Frontiers in microbiology 5, 587, doi:10.3389/fmicb.2014.00587 (2014).

3 Sommer, F. & Bäckhed, F. The gut microbiota—masters of host development and

physiology. Nature Reviews Microbiology 11, 227-238 (2013).

4 Clemente, J. C., Ursell, L. K., Parfrey, L. W. & Knight, R. The impact of the gut microbiota

on human health: an integrative view. Cell 148, 1258-1270 (2012).

5 Nicholson, J. K. et al. Host-gut microbiota metabolic interactions. Science 336, 1262-1267

(2012).

6 Zoetendal, E. G., Akkermans, A. D., Akkermans-van Vliet, W. M., de Visser, J. A. G. & de

Vos, W. M. The host genotype affects the bacterial community in the human gastronintestinal

tract. Microbial ecology in health and disease 13, 129-134 (2001).

7 Redford, K. H., Segre, J. A., Salafsky, N., del Rio, C. M. & McAloose, D. Conservation and

the microbiome. Conserv. Biol. 26, 195-197 (2012).

8 Ley, R. E., Peterson, D. A. & Gordon, J. I. Ecological and evolutionary forces shaping

microbial diversity in the human intestine. Cell 124, 837-848 (2006).

9 Degnan, P. H. et al. Factors associated with the diversification of the gut microbial

communities within chimpanzees from Gombe National Park. Proceedings of the National

Academy of Sciences 109, 13034-13039 (2012).

10 Tung, J. et al. Social networks predict gut microbiome composition in wild baboons. Elife 4,

e05224 (2015).

11 Gomez, A. et al. Gut Microbiome of Coexisting BaAka Pygmies and Bantu Reflects

Gradients of Traditional Subsistence Patterns. Cell reports 14, 2142-2153,

doi:10.1016/j.celrep.2016.02.013 (2016).

12 Barelli, C. et al. Habitat fragmentation is associated to gut microbiota diversity of an

endangered primate: implications for conservation. Scientific reports 5, 14862,

doi:10.1038/srep14862 (2015).

13 Amato, K. R. et al. Habitat degradation impacts black howler monkey (Alouatta pigra)

gastrointestinal microbiomes. The ISME journal 7, 1344-1353, doi:10.1038/ismej.2013.16

(2013).

14 Schwab, C., Cristescu, B., Northrup, J., Stenhouse, G. & Gänzle, M. Diet and environment

shape fecal bacterial microbiota composition and enteric pathogen load of grizzly bears. PLoS

One 6, e27905 (2011).

15 Gomez, A. et al. Gut microbiome composition and metabolomic profiles of wild western

lowland gorillas (Gorilla gorilla gorilla) reflect host ecology. Molecular ecology 24, 2551-

2565 (2015).

16 Phillips, C. D. et al. Microbiome analysis among bats describes influences of host phylogeny,

life history, physiology and geography. Molecular ecology 21, 2617-2627 (2012).

17 Menke, S. et al. Oligotyping reveals differences between gut microbiomes of free-ranging

sympatric Namibian carnivores (Acinonyx jubatus, Canis mesomelas) on a bacterial species-

like level. Frontiers in microbiology 5, 526, doi:10.3389/fmicb.2014.00526 (2014).

18 Amato, K. Co-evolution in context: the importance of studying gut microbiomes in wild

animals. Microbiome Science and Medicine 1 (2013).

19 Podolsky, D. K. The current future understanding of inflammatory bowel disease. Best

practice & research Clinical gastroenterology 16, 933-943 (2002).

20 Targan, S. R. & Karp, L. C. Defects in mucosal immunity leading to ulcerative colitis.

Immunological reviews 206, 296-305 (2005).

21 Round, J. L. & Mazmanian, S. K. The gut microbiota shapes intestinal immune responses

during health and disease. Nature Reviews Immunology 9, 313-323 (2009).

22 Suchodolski, J. S. Intestinal microbiota of dogs and cats: a bigger world than we thought.

Veterinary Clinics of North America: Small Animal Practice 41, 261-272 (2011).

23 Sekirov, I., Russell, S. L., Antunes, L. C. M. & Finlay, B. B. Gut microbiota in health and

disease. Physiological reviews 90, 859-904 (2010).

24 Bahrndorff, S., Alemu, T., Alemneh, T. & Lund Nielsen, J. The microbiome of animals:

implications for conservation biology. International journal of genomics 2016 (2016).

25 Stumpf, R. et al. Microbiomes, metagenomics, and primate conservation: New strategies,

tools, and applications. Biological Conservation 199, 56-66 (2016).

26 Seidensticker, J. Saving wild tigers: a case study in biodiversity loss and challenges to be met

for recovery beyond 2010. Integrative zoology 5, 285-299, doi:10.1111/j.1749-

4877.2010.00214.x (2010).

27 Dinerstein, E. et al. The fate of wild tigers. . BioScience 57, 508-514 (2007).

28 Barber‐Meyer, S. M. et al. Influence of prey depletion and human disturbance on tiger

occupancy in Nepal. Journal of Zoology 289, 10-18 (2013).

29 Kenney, J. S., Smith, J. L., Starfield, A. M. & McDougal, C. W. The long‐term effects of tiger

poaching on population viability. . Conservation Biology. 1995 Oct 1;9(5):1127-33. 9, 1127-

1133 (1995).

30 Smith, J. L., Ahern, S. C. & McDougal, C. Landscape analysis of tiger distribution and

habitat quality in Nepal. . Conservation Biology. 1998 Dec 1;12(6):1338-46. 12, 1338-1346

(1998).

31 Wikramanayake, E. et al. Designing a conservation landscape for tigers in human‐dominated

environments. . Conservation Biology 18, 839-844 (2004).

32 Kanagaraj, R., Wiegand, T., Kramer‐Schadt, S., Anwar, M. & Goyal, S. P. Assessing habitat

suitability for tiger in the fragmented Terai Arc Landscape of India and Nepal. . Ecography.

2011 Dec 1;34(6):970-81. 34, 970-981 (2011).

33 Treves, A. a. K. U. K. Human-carnivore conflict and perspectives on carnivore management

worldwide. Conservation Biology 17, 1491-1499 (2003).

34 Zimmermann, A. et al. Contemporary views of human–carnivore conflicts on wild

rangelands. Wiley-Blackwell, Oxford. (2010).

35 Clayton, J. B. et al. Captivity humanizes the primate microbiome. Proceedings of the

National Academy of Sciences, 201521835 (2016).

36 NTGP/CMDN. Nepal Tiger Genome Project - Final Report 2014, Center for Molecular

Dynamics-Nepal. (Center for Molecular Dynamics-Nepal, 2014).

37 Becker, A. A., Janssens, G. P., Snauwaert, C., Hesta, M. & Huys, G. Integrated community

profiling indicates long-term temporal stability of the predominant faecal microbiota in

captive cheetahs. PLoS One 10, e0123933, doi:10.1371/journal.pone.0123933 (2015).

38 Zhang, H., Liu, G., Chen, L. & Sha, W. Composition and diversity of the bacterial community

in snow leopard (Uncia uncia) distal gut. Annals of Microbiology 65, 703-711 (2015).

39 Handl, S., Dowd, S., Garcia-Mazcorro, J., Steiner, J. & Suchodolski, J. Massive parallel 16S

rRNA gene pyrosequencing reveals highly diverse fecal bacterial and fungal communities in

healthy dogs and cats. FEMS microbiology ecology 76, 301-310 (2011).

40 Chao, A. Nonparametric estimation of the number of classes in a population. Scandinavian

Journal of statistics, 265-270 (1984).

41 Chao, A. Species estimation and applications. Encyclopedia of statistical sciences (2005).

42 Shannon, C. & Weaver, W. (University of Illinois Press IL, 1949).

43 Simpson, E. Measurement of diversity. Nature (1949).

44 Fisher, R. A., Corbet, A. S. & Williams, C. B. The relation between the number of species

and the number of individuals in a random sample of an animal population. The Journal of

Animal Ecology, 42-58 (1943).

45 Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: a practical and powerful

approach to multiple testing. Journal of the Royal Statistical Society. Series B

(Methodological), 289-300 (1995).

46 Woese, C. R. Bacterial evolution. Microbiological reviews 51, 221 (1987).

47 Eaton, K. et al. Gastric spiral bacilli in captive cheetahs. Scandinavian Journal of

Gastroenterology 26, 38-42 (1991).

48 Minamoto, Y., Dhanani, N., Markel, M. E., Steiner, J. M. & Suchodolski, J. S. Prevalence of

Clostridium perfringens, Clostridium perfringens enterotoxin and dysbiosis in fecal samples

of dogs with diarrhea. Veterinary microbiology 174, 463-473 (2014).

49 Frank, D. N. et al. Molecular-phylogenetic characterization of microbial community

imbalances in human inflammatory bowel diseases. Proceedings of the National Academy of

Sciences 104, 13780-13785 (2007).

50 Suchodolski, J. S. et al. The fecal microbiome in cats with diarrhea. PloS one 10, e0127378

(2015).

51 Shin, N.-R., Whon, T. W. & Bae, J.-W. Proteobacteria: microbial signature of dysbiosis in gut

microbiota. Trends in biotechnology 33, 496-503 (2015).

52 Ramakrishna, B. S. Role of the gut microbiota in human nutrition and metabolism. Journal of

gastroenterology and hepatology 28, 9-17 (2013).

53 Kohl, K. D., Sadowska, E. T., Rudolf, A. M., Dearing, M. D. & Koteja, P. Experimental

evolution on a wild mammal species results in modifications of gut microbial communities.

Frontiers in microbiology 7 (2016).

54 Menke, S. et al. Gut microbiomes of free‐ranging and captive Namibian cheetahs: diversity,

putative functions, and occurrence of potential pathogens. Molecular Ecology (2017).

55 Delport, T. C., Power, M. L., Harcourt, R. G., Webster, K. N. & Tetu, S. G. Colony location

and captivity influence the gut microbial community composition of the Australian sea lion

(Neophoca cinerea). Applied and environmental microbiology 82, 3440-3449 (2016).

56 Yildirim, S. et al. Characterization of the fecal microbiome from non-human wild primates

reveals species specific microbial communities. PloS one 5, e13963 (2010).

57 Medeiros, A. W. et al. Characterization of the faecal bacterial community of wild young

South American (Arctocephalus australis) and Subantarctic fur seals (Arctocephalus

tropicalis). FEMS Microbiol Ecol 92, doi:10.1093/femsec/fiw029 (2016).

58 Krajmalnik-Brown, R., Ilhan, Z.-E., Kang, D.-W. & DiBaise, J. K. Effects of gut microbes on

nutrient absorption and energy regulation. Nutrition in Clinical Practice 27, 201-214 (2012).

59 Ismail, N. A. et al. Frequency of Firmicutes and Bacteroidetes in gut microbiota in obese and

normal weight Egyptian children and adults. Archives of medical science: AMS 7, 501 (2011).

60 Boulangé, C. L., Neves, A. L., Chilloux, J., Nicholson, J. K. & Dumas, M.-E. Impact of the

gut microbiota on inflammation, obesity, and metabolic disease. Genome medicine 8, 42

(2016).

61 Maynard, C. L., Elson, C. O., Hatton, R. D. & Weaver, C. T. Reciprocal interactions of the

intestinal microbiota and immune system. Nature 489, 231-241 (2012).

62 Bayhan, G. İ., Tanır, G., Karaman, İ. & Özkan, Ş. Comamonas testosteroni: an unusual

bacteria associated with acute appendicitis. Balkan medical journal 30, 447 (2013).

63 Almuzara, M. N. et al. Intra-abdominal infections due to Comamonas kerstersii. Journal of

clinical microbiology 51, 1998-2000 (2013).

64 Kageyama, A., Sakamoto, M. & Benno, Y. Rapid identification and quantification of

Collinsella aerofaciens using PCR. FEMS microbiology letters 183, 43-47 (2000).

65 Kassinen, A. et al. The fecal microbiota of irritable bowel syndrome patients differs

significantly from that of healthy subjects. Gastroenterology 133, 24-33 (2007).

66 Chen, J. et al. An expansion of rare lineage intestinal microbes characterizes rheumatoid

arthritis. Genome Med 8, 43, doi:10.1186/s13073-016-0299-7 (2016).

67 Kostic, A. D. et al. Fusobacterium nucleatum potentiates intestinal tumorigenesis and

modulates the tumor-immune microenvironment. Cell host & microbe 14, 207-215,

doi:10.1016/j.chom.2013.07.007 (2013).

68 Dulal, S. & Keku, T. O. Gut microbiome and colorectal adenomas. Cancer journal 20, 225-

231, doi:10.1097/PPO.0000000000000050 (2014).

69 Langworth, B. F. Fusobacterium necrophorum: its characteristics and role as an animal

pathogen. Bacteriological Reviews 41, 373 (1977).

70 Edwards, J., Davis, D., Roffe, T., Ramiro-Ibanez, F. & Elzer, P. Fusobacteriosis in captive

wild-caught pronghorns (Antilocapra americana). Veterinary pathology 38, 549-552 (2001).

71 Smith, G., Wallace, L. & Noakes, D. Experimental observations on the pathogenesis of

necrobacillosis. Epidemiology & Infection 104, 73-78 (1990).

72 Rubinstein, M. R. et al. Fusobacterium nucleatum promotes colorectal carcinogenesis by

modulating E-cadherin/β-catenin signaling via its FadA adhesin. Cell host & microbe 14,

195-206 (2013).

73 Seutin, G., White, B. N. & Boag, P. T. Preservation of avian blood and tissue samples for

DNA analyses. Canadian Journal of Zoology 69, 82-90, doi:10.1139/z91-013 (1991).

74 Wultsch, C., Waits, L., Hallerman, E. & Kelly, M. Optimizing Collection Methods for

Noninvasive Genetic Sampling of Neotropical Felids. Wildlife Society Bulletin 39, 403-412

(2015).

75 Bhagavatula, J. & Singh, L. Genotyping faecal samples of Bengal tiger Panthera tigris tigris

for population estimation: a pilot study. BMC genetics 7, 48, doi:10.1186/1471-2156-7-48

(2006).

76 Menotti-Raymond, M. et al. A genetic linkage map of microsatellites in the domestic cat

(Felis catus). Genomics 57, 9-23 (1999).

77 Sharma, R. et al. Fourteen new di‐and tetranucleotide microsatellite loci for the critically

endangered Indian tiger (Panthera tigris tigris). Molecular ecology resources 8, 1480-1482

(2008).

78 Menotti-Raymond, M. A., David, V. A., Wachter, L. L., Butler, J. M. & O‘Brien, S. J. An

STR forensic typing system for genetic individualization of domestic cat (Felis catus)

samples. J Forensic Sci 50, 1061-1070 (2005).

79 Caporaso, J. et al. Global patterns of 16S rRNA diversity at a depth of millions of sequences

per sample. Proceedings of the National Academy of Sciences 108, 4516-4522 (2011).

80 Caporaso, J. et al. QIIME allows analysis of high-throughput community sequencing data.

Nature methods 7, 335-336 (2010).

81 DeSantis, T. et al. Greengenes, a chimera-checked 16S rRNA gene database and workbench

compatible with ARB. Applied and environmental microbiology 72, 5069-5072 (2006).

82 Caporaso, J. et al. PyNAST: a flexible tool for aligning sequences to a template alignment.

Bioinformatics 26, 266-267 (2010).

83 Price, M., Dehal, P. & Arkin, A. FastTree 2–approximately maximum-likelihood trees for

large alignments. PloS one 5, e9490 (2010).

84 RDCT. R Development Core Team (2014). R: a language and environment for statistical

computing (R Foundation for Statistical Computing). (2014).

85 McMurdie, P. J. & Holmes, S. phyloseq: an R package for reproducible interactive analysis

and graphics of microbiome census data. PloS one 8, e61217 (2013).

86 Anderson, M. J. A new method for non-parametric multivariate analysis of variance. Austral

Ecology 26, 32-46, doi:10.1111/j.1442-9993.2001.01070.pp.x (2001).

87 Anderson, M. J. & Willis, T. J. Canonical analysis of principal coordinates: a useful method

of constrained ordination for ecology. Ecology 84, 511-525 (2003).

88 Anderson, M. J., Ellingsen, K. E. & McArdle, B. H. Multivariate dispersion as a measure of

beta diversity. Ecology letters 9, 683-693 (2006).

89 Anderson, M., Gorley, R. N. & Clarke, R. K. Permanova+ for Primer: Guide to Software and

Statisticl Methods. (Primer-E Limited, 2008).

90 Langille, M. G. et al. Predictive functional profiling of microbial communities using 16S

rRNA marker gene sequences. Nature biotechnology 31, 814-821 (2013).

91 Markowitz, V. M. et al. IMG: the integrated microbial genomes database and comparative

analysis system. Nucleic acids research 40, D115-D122 (2011).

92 Kanehisa, M., Goto, S., Sato, Y., Furumichi, M. & Tanabe, M. KEGG for integration and

interpretation of large-scale molecular data sets. Nucleic acids research 40, D109-D114

(2011).

93 Parks, D. H., Tyson, G. W., Hugenholtz, P. & Beiko, R. G. STAMP: statistical analysis of

taxonomic and functional profiles. Bioinformatics 30, 3123-3124 (2014).

94 Gower, J. C. Generalized procrustes analysis. Psychometrika 40, 33-51 (1975).

CHAPTER VII

Conclusion and future direction

The results of my study provide a holistic perspective on conservation research of Bengal tiger in

particular and its symbiotic biodiversity in general using non-invasive genetics tools and technology.

Understanding general biodiversity while focusing on flagship species:

More research is needed to understand the type of interaction between the carnivore species existing

in the Terai and trans-himalaya landscapes. My study helps our understanding of other sympatric

species prevalent in the tiger habitat. As tiger is a flagship species, it often attracts both attention and

resources, leveraging on that, I have included efforts to also incorporate DNA analysis of samples

that were not found to be of tiger and by doing so I was able to create a biodiversity map of the terai

landscape. Biodiversity in general is not well studied in this part of the world, and I believe my study

has provided an opportunity to create baseline a biodiversity map of sympatric species in the tiger

habitat. Additionally, my analysis has also identified the presence of an unknown species of fox in the

Terai landscape. DNA sequence of identified species closely matched with the red fox (98% match;

100% query on 90bp PCR amplicon DNA sequence data) but red foxes are not known to occur in the

Terai landscape. We believe that we have detected Bengal fox (Vulpes bengalensis), which is known

to inhabit the Terai area of Nepal and India, but there were no DNA sequences for Bengal fox in the

NCBI database, so further study is needed to verify our assumption.

Identifying factors that impact species (tiger) conservation status:

High rates of anthropogenic transformation of landscapes play a major role in extinction of wild

mammal populations. This kind of land-use change has been devastating for predators like tiger with

the loss of habitat across their historical distribution, which often results in severe population decline

and local extirpations in the last few generations. Most wild tiger populations are now confined to

protected areas. Poaching and habitat loss are major threats to the tiger and other wildlife species in

the remaining 13,500 km2 of potential habitat in Nepal. Against this background, the results of my

genetic study of tigers are critical to understanding the current levels of genetic diversity and

connectivity which can play a significant role in prioritizing conservation efforts in the TAL species

like tigers.

Tigers in the TAL displayed moderate levels of genetic variation (He = 0.61) across the landscape,

perhaps due to moderate population size and/or gene flow. Overall, my results suggest low to

moderate levels of genetic diversity based on eight polymorphic loci across the tiger population in the

TAL. Lower levels of genetic diversity were detected in this population compared to other high-

diversity areas in the Satpura-Maikal landscape of India. Results suggest three distinct clusters of tiger

sub-populations within the landscape. Admixed models and concurrent results from both spatial and

non-spatial algorithms in the STRUCTURE and TESS programs suggest three (k = 3) clusters of tiger

sub-populations in the landscape, confirming my prior hypothesis. The field data suggested

demographic contiguity in and around the protected areas. However, in the absence of tiger samples

from the Indian side of the TAL, there is a lack of clarity in explaining the overall genetic structure of

tigers in the entire landscape; tigers could travel through the contiguous protected areas of India into

Nepal.

Tiger’s greatest threat- Poaching: developing of utility based research study

Wildlife crime is a serious transnational threat to biodiversity in South Asia. Driven by organized

criminal syndicates that control the burgeoning and highly lucrative illicit trade in wildlife parts, there

is an immediate need to develop tools that can assist in tracking and ultimately preventing wildlife

crime. The use of forensic tools plays an important role in strengthening law enforcement efforts to

address wildlife crime and biodiversity conservation. It has already been successfully used in some

countries to improve investigation of wildlife cases, identification of poaching hotspots and

trafficking routes as well as prosecution of wildlife criminals by connecting individuals to crime

scenes.

Application of molecular (DNA) forensics is relatively new and the rapidly evolving technology is

gaining acceptance in wildlife conservation efforts. The information obtained through the genetic-

forensic approach is more robust and comprehensive than conventional methods. To fulfill the urgent

need to address location of poached tigers, I created a wild tiger geo-spatial, genetic baseline database

through the Nepal Tiger Genome Project (2011-2013) which is probably the world's most

comprehensive genetic database. This has provided valuable information on wildlife crime ―hot spots‖

and presented scientific evidence to incriminate criminals in a court of law. I identified one tiger skin

that had an exact DNA profile with one of our reference NTGP tigers from BNP. I collected a fecal

sample of that individual in the winter season of 2012; and since I collected fecal samples that were

relatively fresh (<3 month), this particular individual was probably poached as recent as 1 to 2 years

ago. Of all the forensic evidence I analyzed, most were traced to Bardia, highlighting this area as a

wildlife crime hotspot.

This was crucial information for Nepal‘s law enforcement officials. After dissemination of these

results, the concerned stakeholders, including the Nepal Police and the Nepal Army increased their

patrols and intelligence activities particularly in the Bardia region. As a result, a few criminal

networks involved in wildlife crime in that area were discovered and necessary actions were taken

supporting long term wildlife crime prevention.

Comparative genomics study on host (tiger) and its gut micobiome:

The gut microbial composition obtained with fecal samples collected in the tiger sub-population

showed that while there is a core microbiome common to all the three subpopulations, the relative

abundance of various taxa is significantly different. These differences are highest between

subpopulations of CNP and SWR. This inference corresponds to the geo-specific locations of these

regions. However additionally, predicted metabolic functionalities also reveal a consistent pattern

with higher functional enrichment and diversity in CNP for most categories and the lowest enrichment

and diversity in SWR. The enrichment distribution based on samples from BNP overlap at two

extremes with the CNP and SWR based enrichment distributions. My results map the genetic

diversity, gut microbiome composition and predictive metabolic functionalities based on fecal

samples collected over the Terai region in Nepal. I find a consistent pattern in differentiation, gut

microbial composition and predicted functionality that corresponds to habitat fragmentation. My

findings highlight the necessity of a comprehensive systems biology based approach to conservation

with a focus on monitoring and maintaining genetic diversity and a healthy gut microbial

composition.

Overall implication, utility and future direction of my work:

My study, through Nepal Tiger Genome Project, has looked into all aspect of molecular approaches

and created the most comprehensive genetic baseline database of wild tigers in Nepal and helped

comprehend the tiger population from conservation genetics perspective. Information uncovered has

now been utilized to formulate effective conservation strategies. This kind of work and effort is

carried out for the first time in Nepal; and some work, especially molecular forensics and gut

microbiome study, are done for the first time. This has also created necessary capacity to do similar

work in other important endangered species in developing countries like Nepal. Additionally I am

also looking at creating viral and dietary profiles on some of these tiger fecal samples as part of my

commitment to understand this species further. Meanwhile, I have been able to train over 50

conservation biologists and laboratory specialists on various aspects of our study.

~~ END of REPORT~~