legacy of draught cattle breeds of south india: insights into … · 2021. 1. 21. · 2 35 legacy...

58
1 1 Title : Legacy of draught cattle breeds of South India: Insights into population 2 structure, genetic admixture and maternal origin 3 4 Authors : Vandana Manomohan 1,2 , Ramasamy Saravanan 2 , Rudolf Pichler 1 , Nagarajan 5 Murali 2 , Karuppusamy Sivakumar 2 , Krovvidi Sudhakar 3 , Raja K 6 Nachiappan 4 and Kathiravan Periasamy 1,5* 7 8 Affiliations : 1 Animal Production and Health Section, Joint FAO/IAEA Division of 9 Nuclear Techniques in Food and Agriculture, International Atomic Energy 10 Agency, Vienna, Austria 11 2 Veterinary College and Research Institute, Namakkal, Tamil Nadu 12 Veterinary and Animal Sciences University, Chennai, India; 13 3 NTR College of Veterinary Science, Gannavaram, Sri Venkateswara 14 Veterinary University, Tirupati, Andhra Pradesh, India; 15 4 National Bureau of Animal Genetic Resources, Karnal, Haryana, India; 16 5 Animal Genetics Resources Branch, Animal Production and Health 17 Division, Food and Agriculture Organization of the United Nations, Rome, 18 Italy. 19 20 Corresponding: Kathiravan Periasamy 21 author 22 23 24 Address : Animal Production and Health Laboratory, Joint FAO/IAEA Division of 25 Nuclear Techniques in Food and Agriculture, International Atomic Energy 26 Agency, Seibersdorf, Vienna, Austria 27 28 29 Email : [email protected]; [email protected] 30 31 32 Telephone : 00431260027328 33 34 . CC-BY 4.0 International license perpetuity. It is made available under a preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in The copyright holder for this this version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560 doi: bioRxiv preprint

Upload: others

Post on 05-Feb-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • 1

    1 Title : Legacy of draught cattle breeds of South India: Insights into population

    2 structure, genetic admixture and maternal origin

    3

    4 Authors : Vandana Manomohan1,2, Ramasamy Saravanan 2, Rudolf Pichler1, Nagarajan

    5 Murali2, Karuppusamy Sivakumar2, Krovvidi Sudhakar3, Raja K

    6 Nachiappan4 and Kathiravan Periasamy1,5*

    78 Affiliations : 1Animal Production and Health Section, Joint FAO/IAEA Division of

    9 Nuclear Techniques in Food and Agriculture, International Atomic Energy

    10 Agency, Vienna, Austria

    11 2Veterinary College and Research Institute, Namakkal, Tamil Nadu

    12 Veterinary and Animal Sciences University, Chennai, India;

    13 3NTR College of Veterinary Science, Gannavaram, Sri Venkateswara

    14 Veterinary University, Tirupati, Andhra Pradesh, India;

    15 4National Bureau of Animal Genetic Resources, Karnal, Haryana, India;

    16 5Animal Genetics Resources Branch, Animal Production and Health

    17 Division, Food and Agriculture Organization of the United Nations, Rome,

    18 Italy.

    19

    20 Corresponding: Kathiravan Periasamy21 author22

    23

    24 Address : Animal Production and Health Laboratory, Joint FAO/IAEA Division of

    25 Nuclear Techniques in Food and Agriculture, International Atomic Energy

    26 Agency, Seibersdorf, Vienna, Austria

    27

    28

    29 Email : [email protected]; [email protected]

    30

    31

    32 Telephone : 00431260027328

    33

    34

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 2

    35 Legacy of draught cattle breeds of South India: Insights into population structure,

    36 genetic admixture and maternal origin

    37

    38 Vandana Manomohan1,2, Ramasamy Saravanan2, Rudolf Pichler1, Nagarajan Murali2,

    39 Karuppusamy Sivakumar2, Krovvidi Sudhakar3, Raja K Nachiappan4

    40 and Kathiravan Periasamy1,5*

    41 1Animal Production and Health Section, Joint FAO/IAEA Division, International Atomic

    42 Energy Agency, Vienna, Austria; 2Veterinary College and Research Institute, Namakkal,

    43 Tamil Nadu Veterinary and Animal Sciences University, Chennai, India; 3NTR College of

    44 Veterinary Science, Gannavaram, Sri Venkateswara Veterinary University, Tirupati, Andhra

    45 Pradesh, India; 4National Bureau of Animal Genetic Resources, Karnal, Haryana, India;

    46 5Animal Genetics Resources Branch, Animal Production and Health Division, Food and

    47 Agriculture Organization of the United Nations, Rome, Italy.

    48 *Corresponding author email: [email protected]; [email protected] 49

    50

    51 ABSTRACT

    52 The present study is the first comprehensive report on diversity, population structure, genetic

    53 admixture and mitochondrial DNA variation in South Indian draught type zebu cattle. The

    54 diversity of South Indian cattle was moderately higher. A significantly strong negative

    55 correlation coefficient of -0.674 (P6.25%) was observed in Punganur, Vechur, Umblachery and Pulikulam cattle breeds. Two

    63 major maternal haplogroups, I1 and I2, typical of zebu cattle were observed, with the former

    64 being predominant than the later. The pairwise differences among the I2 haplotypes of South

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 3

    65 Indian cattle were relatively higher than West Indian (Indus valley site) zebu cattle. The results

    66 indicated the need for additional sampling and comprehensive analysis of mtDNA control

    67 region variations to unravel the probable location of origin and domestication of I2 zebu

    68 lineage. The present study also revealed major concerns on South Indian zebu cattle (i) risk of

    69 endangerment due to small effective population size and high rate of inbreeding (ii) lack of

    70 sufficient purebred zebu bulls for breeding and (iii) increasing level of taurine admixture in

    71 zebu cattle. Availability of purebred semen for artificial insemination, incorporation of

    72 genomic/molecular information to identify purebred animals and increased awareness among

    73 farmers will help to maintain breed purity, conserve and improve these important draught cattle

    74 germplasms of South India.

    75

    76 Key words: Zebu, diversity, inbreeding, genetic relationship, taurine introgression,

    77 haplogroup, mismatch distribution, domestication

    78

    79 1. INTRODUCTION

    80 Ever since domestication, animals have been used as an important source of draught power for

    81 agricultural production. Among them, cattle have been used for a variety of farm operations

    82 like ploughing, carting, tilling, sowing, weeding, water lifting, threshing, oil extraction,

    83 sugarcane crushing and transport. Over time, with increasing mechanization of agriculture, the

    84 use of draught cattle in farm operations reduced significantly and became negligible in

    85 advanced economies. However, in developing countries, particularly in small holder

    86 production systems and in hilly and mountainous regions, agricultural production still relies on

    87 draught animal power (Zhou et al. 2018). For example, 55% of the smallholder farmers in

    88 Swaziland rely on draft animal power for land cultivation, and more than 88% of draft animals

    89 found on the Swazi National Land are cattle (Kienzle et al. 2013). In India, the significance of

    90 draught animal power has reduced greatly in the last few decades with its share in total farm

    91 power availability declining from about 78% in 1960-61 to 7% in 2013-2014 (Singh et al. 2014)

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 4

    92 and expected to further reduce to around 5% by 2020 (Natarajan et al. 2016). Over the years,

    93 the annual use of draught animals declined from about 1200-1800 hours to 300-500 hours per

    94 pair per year. Consequently, the population of draught cattle steadily decreased at a negative

    95 annual compound growth rate of about -0.8%, resulting in significant decline of purebred native

    96 breeds.

    97 India is home to the largest cattle population (13.1 % of world’s cattle population) in the world

    98 which constitutes 37.3 % of its total livestock. As per the 19th livestock census, India has 190.9

    99 million cattle, out of which 151 million is indigenous cattle. The diversity of Indian cattle is

    100 represented by 60 local breeds, eight regional transboundary breeds and seven international

    101 transboundary breeds (FAO, 2015). Of these, 41 have been registered by National Bureau of

    102 Animal Genetic Resources, the nodal agency for the registration of livestock breeds in India.

    103 The Indian Zebu cattle (Bos indicus) breeds have been classified into three major groups: dairy

    104 type, draught type and dual-purpose cattle. The dairy type and dual-purpose cattle breeds are

    105 predominantly distributed in North and North-Western India, while most indigenous draught

    106 type cattle breeds are located in Southern and Eastern India. Apart from rapid mechanization

    107 of agriculture in South India, the increasing cost of maintaining draught type cattle forced the

    108 farmers to replace these animals with crossbred cattle that fetched better returns from the sale

    109 of milk to dairy cooperatives. During late 1980s and early 1990s, the draught cattle breeds

    110 started losing their economic relevance and the farmers increasingly resorted to crossbreeding

    111 with commercial taurine cattle like Jersey, Holstein-Friesian and Brown Swiss. Such crossbred

    112 cows approximately yielded more than double the milk as that of indigenous draught breeds.

    113 Combined with farmers’ preference for crossbreds, the relatively easy access to breeding

    114 services (through state animal husbandry departments, state cooperatives and private

    115 inseminators, etc.) and availability of better infrastructure for artificial insemination and

    116 veterinary care, the proportion of crossbred cattle among the breedable females became much

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 5

    117 higher in South Indian states than rest of India. The crossbred cattle constitute 35.9% of total

    118 cattle in Andhra Pradesh, 41.9% in Karnataka, 76.1% in Tamil Nadu and 94.6% in Kerala as

    119 against the national average of 27.7%. This resulted in significant genetic dilution of draught

    120 cattle breeds of South India with varying levels of zebu-taurine admixture.

    121 Indiscriminate crossbreeding, lack of stabilization of exotic inheritance and absence of

    122 selection among F2 hybrids led to negative impacts like increased disease susceptibility,

    123 decreased fertility and drop in production levels among crossbred cattle. During the last two

    124 decades, there has been a renewed interest in the conservation of purebred indigenous draught

    125 cattle breeds and implementation of selective breeding programs targeting a range of traits.

    126 Selective breeding schemes among certain purebred native breeds like Kangayam, Ongole,

    127 Deoni and Hallikar were put in place by making available the purebred frozen semen for

    128 artificial insemination. Such programs have been constrained by the availability of superior

    129 genetic merit bulls and unavailability of sufficient of number of true to the type, purebred bulls.

    130 With the absence of pedigree records, the level of taurine inheritance in most purebred draught

    131 cattle populations is not clearly understood. Further, information on genetic differentiation,

    132 population structure and maternal inheritance among South Indian cattle breeds is scanty.

    133 Hence, the present study was undertaken with the following objectives: (i) assess biodiversity

    134 and population structure among purebred South Indian draught cattle breeds (ii) estimate level

    135 of genetic admixture among purebred zebu cattle and (iii) identify mitochondrial DNA based

    136 maternal lineages and evaluate phylogeography of South Indian cattle.

    137

    138 2. MATERIALS AND METHODS

    139 2.1. Animal ethics statement

    140 Blood samples were collected by jugular venipuncture into EDTA vacutainer tubes from

    141 various purebred zebu, commercial taurine and crossbred cattle. Blood sampling was

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 6

    142 performed by local veterinarians in the respective native breed tracts following the standard

    143 good animal practice. The consent was obtained from animal owners for the usage of the

    144 samples in this study. Therefore, no further license from the “Institutional Committee for Care

    145 and Use of Experimental Animals” of the Tamil Nadu Veterinary and Animal Sciences

    146 University, Chennai, India and the Joint FAO/IAEA Division, International Atomic Energy

    147 Agency, Vienna, Austria was required.

    148 2.2. Sample collection

    149 A total of 529 blood samples were collected from cattle representing 14 breeds/populations

    150 that are located in South/Peninsular Indian states: Ongole (n=51) and Punganur (n=18) from

    151 Andhra Pradesh; Hallikar (n=36) and Deoni (n=53) from Karnataka; Vechur (n=26) from

    152 Kerala; Alambadi (n=29), Bargur (n=52), Kangayam (n=51), Pulikulam (n=34), Umblachery

    153 (n=33), Holstein-Friesian (n=15), Jersey (n=34), Holstein-Friesian crossbred (n=39) and Jersey

    154 crossbred (n=58) from Tamil Nadu. A stratified random sampling procedure was followed to

    155 collect blood from unrelated cattle with typical phenotypic features and located in randomly

    156 selected villages of the native breed tract. With the absence of pedigree records in most

    157 instances, unrelatedness was ensured by interviewing the farmers about the breeding history.

    158 Genomic DNA was extracted by standard phenol chloroform method (Sambrook and Russell,

    159 2001). The DNA samples were subjected to quality control using Nanodrop spectrophotometer

    160 and stored at −20ºC until further processing.

    161 2.3. Microsatellite genotyping

    162 All the DNA samples were subjected to PCR amplification using 27 FAO recommended

    163 microsatellite markers (FAO, 2011; Supplementary Table ST1). The forward primers

    164 conjugated to one of the three fluorescent dyes (FAM, HEX and ATTO550) were used for

    165 genotyping these marker loci. Polymerase chain reaction was performed under following

    166 conditions: initial denaturation for 5 min at 950C, followed by 35 cycles of denaturation for 30s

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 7

    167 at 950C, 1 minute at specific annealing temperatures of marker locus, elongation at 720C for 1

    168 min with the final extension at 720C for 10 min. The PCR products were then electrophoresed

    169 after multiplexing in an automated DNA analyzer ABI3100 (Applied Biosystems, USA) with

    170 ROX500 (Applied Biosystems, USA) as internal lane control. All the 27 STR loci were

    171 multiplexed in six sets for genotyping as shown in Supplementary Table ST1. Determination

    172 of allele size and extraction of genotype data was performed using GENEMAPPER software.

    173 2.4. Sequencing mitochondrial DNA control region

    174 Primers were custom designed and synthesized to amplify 1207 bp bovine mitochondrial DNA

    175 control region using Primer 3 version 4.0 (http://bioinfo.ut.ee/primer3-0.4.0/). The amplified

    176 mtDNA control region included nucleotide positions 15167 to 30 of the complete bovine

    177 mitochondrial genome sequence available at NCBI GenBank accession number AF492350

    178 (Hiendleder et al. 2008). The location and sequence of forward and reverse primers are as

    179 follows: BTMTD1-F (Positions 15167-15186; 5’ AGGACAACCAGTCGAACACC 3’) and

    180 BTMTD1-R 5’ (Positions 11-30; GTGCCTTGCTTTGGGTTAAG 3’). The amplicon included

    181 complete D-loop (control region) flanked by partial cytochrome B, complete t-RNA-Thr and

    182 complete t-RNA-Pro coding sequences at 5’ end while partial t-RNA-Phe sequence flanked the

    183 3’end. Polymerase chain reaction was performed in a total reaction volume of 40µl with the

    184 following cycling conditions: initial denaturation at 95°C for 15 min followed by 30 cycles of

    185 95°C for 1 min; 58°C for 1 min; 72°C for 1 min with final extension at 72°C for 10 min.

    186 Purified PCR products were sequenced using Big Dye Terminator Cycle Sequencing Kit

    187 (Applied Biosystems, U.S.A) on an automated Genetic Analyzer ABI 3100 (Applied

    188 Biosystems, U.S.A).

    189 2.5. Statistical analysis

    190 The presence of null alleles in the dataset was checked using Microchecker version 2.2.3

    191 (Oosterhout et al. 2004). The neutrality of microsatellite markers used in the present study was

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 8

    192 evaluated by FST outlier approach as implemented in LOSITAN version 1 for windows (Antao

    193 et al. 2008). Basic diversity indices (allelic diversity and heterozygosity) and genetic distances

    194 (pairwise Nei’s genetic distance and pairwise inter-individual allele sharing distance) among

    195 South Indian cattle were estimated using MICROSATELLITE ANALYZER (MSA) version

    196 4.05 (Dieringer and Schlotterer, 2003). A UPGMA tree was constructed based on pairwise

    197 Nei’s genetic distances using PHYLIP version 3.5 (Felsenstein, 1993). F statistics for each

    198 locus (Weir and Cockerman, 1984) was calculated and tested using the program FSTAT

    199 version 2.9.3.2. The exact tests for Hardy-Weinberg Equilibrium to evaluate both

    200 heterozygosity deficit and excess at each marker locus in each population (HWE) were

    201 performed using GENEPOP software. Analysis of molecular variance (AMOVA) was

    202 performed using ARLEQUIN version 3.1 (Excoffier et al. 2005). Multidimensional scaling

    203 display of pairwise FST was conducted using SPSS version 13.0. Bayesian clustering analysis

    204 was performed to evaluate population structure and genetic admixture among South Indian

    205 cattle using STRUCTURE 2.3.4 (Pritchard et al. 2000). The program was run with the

    206 assumption of K=2 to K=15 with at least 20 runs for each K. The number of burn in periods

    207 and MCMC repeats used for these runs were 200000 and 200000 respectively. The procedure

    208 of Evanno et al. (2005) was used to estimate the second order rate of change in LnP(D) and

    209 identify the optimal K.

    210 Mitochondrial DNA sequence editing, assembly of contigs and multiple sequence alignments

    211 were performed using Codon Code Aligner version 3.7.1. Mitochondrial DNA diversity

    212 parameters including nucleotide diversity, haplotype diversity, parsimony informative sites and

    213 average number of nucleotide differences were calculated using DnaSP, version 4.10 (Rozas,

    214 2009). Frequency of different haplotypes, haplotype sharing and pairwise FST were estimated

    215 using ARLEQUIN 2.1. MEGA version 6.0 (Tamura et al. 2011) was utilized to identify optimal

    216 DNA substitution model and construct maximum-likelihood phylogeny. HKY+I+G was

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 9

    217 identified as the optimal model for the current sequence dataset based on Akaike information

    218 criterion (AIC). The following sequences were used as references for different maternal

    219 lineages reported in zebu and taurine cattle: I1 (AB085923, AB268563, AB268579,

    220 EF417970, EF417971 and L27733), I2 (AB085921, AB268559, AB268570, AB268573,

    221 AB268581 and AY378133), T1 (DQ124399, EU177841 and EU177842), T2 (DQ124383,

    222 DQ124393, EU177849 and EU177851), T3 (EU177815, EU177816, EU177838 and

    223 EU177839), T4 (DQ124372, DQ124375, DQ124377 and DQ124392), and T5 (EU177862,

    224 EU177863, EU177864 and EU177865). Additionally, 298 mtDNA control region sequences

    225 deposited to GenBank from Bhutan (n=121), Nepal (n=14), China (n=47), North India (n=53),

    226 West India (n=45) and Vietnam (n=18) were utilized for comparative analysis. The accession

    227 number and details of GenBank sequences used in the study are provided in Supplementary

    228 Table ST2. Pairwise FST derived from mtDNA haplotype frequency were utilized to perform

    229 principal components analysis using SPSS version 13.0. Mismatch distribution and the tests

    230 for demographic expansion were performed using ARLEQUIN 3.1. Three parameters viz. tau,

    231 theta 0 and theta 1 were estimated under the sudden expansion model and the tests for goodness

    232 of fit were performed based on sum of square deviations (SSD) between the observed and

    233 expected mismatch and Harpending’s raggedness index. The South Indian cattle breeds were

    234 subjected to tests for selective neutrality through estimation of Tajima’s D and Fu’s FS. Median

    235 Joining (MJ) network of South Indian cattle haplotypes was reconstructed using NETWORK

    236 4.5.1.2 (Bandelt et al. 1999).

    237

    238 3. RESULTS

    239 3.1. Basic diversity measures and Hardy-Weinberg equilibrium

    240 In the present study, a total of 13932 genotypes were generated across 27 microsatellite loci in

    241 516 cattle belonging to 10 draught type zebu breeds, 2 commercial taurine breeds and 2

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 10

    242 crossbred populations. The mean basic diversity indices of different South Indian cattle breeds

    243 are presented in Table 1. Allelic diversity in terms of mean observed number of alleles varied

    244 from 5.93 (Punganur) to 8.07 (Pulikulam) across different zebu cattle breeds, while it was 4.7

    245 and 5.81 in the commercial taurine breeds, Jersey and Holstein-Friesian respectively. The

    246 observed heterozygosity ranged from 0.598 (Kangayam) to 0.671 (Alambadi and Punganur)

    247 while the expected heterozygosity varied between 0.637 (Kangayam) and 0.727 (Punganur)

    248 among the South Indian zebu cattle. Similarly, allelic richness was lowest in Kangayam and

    249 highest in Pulikulam cattle. Assessment of within breed diversity based on inter-individual

    250 allele sharing distance revealed Kangayam with the lowest mean distance among individuals

    251 (0.568) while Pulikulam cattle showed the highest mean distance (0.647). The allelic diversity

    252 observed among South Indian zebu cattle was relatively low as compared to that of West-

    253 Central Indian (Shah et al. 2012), North Indian (Sharma et al. 2015) and East Indian (Sharma

    254 et al. 2013) cattle. The primary domestication centre of zebu (Bos indicus) cattle in North-

    255 Western part of India near Indus valley could explain the higher diversity of North Indian cattle

    256 compared to that of South Indian breeds. Despite the relatively low allelic diversity, the

    257 observed and expected heterozygosity of South Indian cattle were moderately higher (>60%)

    258 and comparable to that of West-Central, North and East Indian cattle (Shah et al. 2012; Sharma

    259 et al. 2013, 2015). As expected, the crossbred cattle (Jersey crossbreds and Holstein-Friesian

    260 crossbreds) exhibited a higher level of allelic diversity and heterozygosity than indigenous zebu

    261 cattle.

    262 The mean estimated FIS values were positive in all the zebu cattle breeds and ranged between

    263 0.027 (Hallikar) and 0.108 (Pulikulam). Breeds like Umblachery, Vechur, Ongole and Bargur

    264 showed mean FIS greater than 7% indicating significant heterozygosity deficit, possibly due to

    265 close breeding and potential inbreeding. The mean FIS estimates observed in most South Indian

    266 cattle breeds were higher than those reported for East Indian (Sharma et al. 2013) and most

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 11

    267 North Indian cattle with the exception of Mewati and Gaolao (Sharma et al. 2015). However,

    268 the mean FIS estimates reported for West-Central Indian cattle were much higher than that of

    269 South Indian cattle (Shah et al. 2012). Similarly, the mean FIS for Vechur and Ongole cattle

    270 observed in the present study was significantly lower than reported previously by Radhika et

    271 al. (2018) and Sharma et al. (2015) respectively. The test for Hardy-Weinberg equilibrium

    272 (HWE) was conducted on a total of 378 breed X locus combinations. Among the 270 breed X

    273 locus combinations tested for zebu cattle, about 62 (22.96%) deviated significantly (P

  • 12

    292 significant outliers with high FST values, indicating positive selection in force (Supplementary

    293 Figure SF1). Of these, HEL5 exhibited significant heterozygosity deficit and consequent

    294 deviation from HWE in 12 out of 14 studied populations. Both the loci have been reported to

    295 be associated with production and/or reproduction traits in cattle (de Oliveira et al. 2005; Li et

    296 al. 2010; Vanessa et al. 2018). Specifically, the homozygous long alleles at HEL5 (allele size

    297 159-169) were significantly associated (p=0.022) with long calving interval in Brangus-Ibage

    298 cattle, a composite breed with taurine and zebu ancestry (de Oliveira et al. 2005). In the present

    299 study, significant difference in the frequency of long (allele size 160-180) and short alleles

    300 (allele size 148-158) was observed among the zebu, taurine and crossbred cattle, resulting in

    301 higher estimation of FST. Selective processes affect neutral loci when the latter is in linkage

    302 disequilibrium with other loci subjected to selection. Considering the possibility of selective

    303 forces operating at the two FST outlier loci, HEL5 and HEL13 were removed from subsequent

    304 analysis of genetic relationship, admixture and population structure of South Indian cattle.

    305 3.2. Genetic differentiation and phylogeny

    306 The global F statistics among the South Indian cattle breeds are presented in Supplementary

    307 Table ST3. The overall global FIT, FST and FIS among zebu, taurine and crossbred cattle was

    308 0.143, 0.101 and 0.047 respectively, while those estimates among zebu cattle alone were 0.109,

    309 0.057 and 0.056 respectively. The results indicated 10.1% of the total genetic variation among

    310 all the studied cattle were due to between breed differences and it reduced to 5.7% when only

    311 zebu cattle breeds were considered. The global FST among all cattle breeds/populations ranged

    312 from 0.052 (INRA005) to 0.223 (ETH152), whereas among zebu cattle breeds, it varied

    313 between 0.024 (INRA005) and 0.114 (HAUT24). Among the investigated loci, ETH152

    314 provided relatively more information on zebu-taurine divide while HAUT24 provided

    315 significant information on differences among zebu cattle breeds. The global FST among South

    316 Indian zebu cattle observed in the present study was comparable to that of East Indian zebu

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 13

    317 (Sharma et al. 2013) cattle. However, much higher genetic differentiation had been reported

    318 among West-Central (Shah et al. 2012) and North Indian (Sharma et al. 2015) cattle breeds.

    319 The pairwise FST and Nei’s genetic distance among the draught type zebu, taurine and crossbred

    320 cattle are presented in Table 2. The pairwise FST ranged from 0.0234 (Hallikar-Alambadi) to

    321 0.107 (Kangayam-Punganur) among the South Indian zebu cattle breeds. Alambadi cattle was

    322 observed to be genetically less differentiated from Hallikar (FST=0.0234), Pulikulam

    323 (FST=0.0237) and Bargur (FST=0.0259) while Kangayam cattle showed highest differentiation

    324 from Punganur (FST=0.107), Vechur (FST=0.1034) and Hallikar (FST=0.0933). Interestingly,

    325 significantly high genetic differentiation (FST=0.0919) was observed among the two

    326 dwarf/miniature type Vechur and Punganur cattle. Similar findings were observed with respect

    327 to the estimates of pairwise Nei’s genetic distance among the South Indian zebu cattle breeds.

    328 The radial tree and dendrogram constructed following UPGMA algorithm based on pairwise

    329 Nei’s genetic distance is presented in Figure 1. Broadly, the taurine, zebu and their crossbred

    330 cattle clustered separately as expected. Among the zebu cattle, a major cluster was formed by

    331 Alambadi, Bargur, Umblachery, Pulikulam, Deoni, Ongole and Hallikar, within which three

    332 sub-clusters were observed. Alambadi and Bargur formed the first sub-cluster, Umblachery

    333 and Pulikulam formed the second sub-cluster while Deoni, Ongole and Hallikar formed the

    334 third sub-cluster. Kangayam, Vechur and Punganur cattle breeds did not form part of any of

    335 these three clusters and were distinct among themselves.

    336 3.3. Genetic relationship and population structure

    337 Understanding genetic relationship and detecting the cryptic genetic structure is an important

    338 first step to elucidate the demography and ancestral history of cattle populations. The

    339 multidimensional scaling (MDS) analysis were used to investigate the relationship among the

    340 studied cattle breeds. Multidimensional scaling (MDS) is an ordination technique used to

    341 display information contained in a distance matrix and to visualize relationship among

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 14

    342 variables within a dataset. This method helps to reduce the dimension of data and to detect

    343 structure among populations in a typical genetic diversity study. The MDS plots derived from

    344 pairwise FST values are presented in Figure 2. The observed s-stress value was 0.01202 when

    345 all the breeds/populations were considered while it was 0.08001 when the zebu cattle alone

    346 were considered. The MDS plots showed a pattern of clustering similar to that of phylogeny.

    347 The taurine and crossbreds clustered separately, while Kangayam, Vechur and Punganur cattle

    348 were distinct from rest of the zebu cattle breeds.

    349 Bayesian clustering of individuals without prior population information was performed to

    350 assess the cryptic genetic structure in the studied cattle breeds. Among K=2 to 15, the dataset

    351 was best described with K=2 genetic clusters at which the second order rate of change in

    352 LnP(D) was maximum (Supplementary Figure SF2). At K=2, the clustering of individuals

    353 followed the zebu-taurine divide (Figure 3). When K=3 was assumed, Zebu cattle breeds got

    354 clustered based on Hallikar or Kangayam ancestry. Both Hallikar and Kangayam cattle were

    355 assigned to their respective clusters with >90% proportion of membership coefficient. Deoni

    356 (80%) and Ongole (69.2%) were predominantly assigned to Kangayam ancestral cluster while

    357 Vechur (86.4%), Punganur (78.7%), Bargur (75.7%) and Alambadi (72.8%) cattle were

    358 predominantly assigned to Hallikar cluster. Pulikulam and Umblachery cattle showed

    359 admixture from both the ancestry. When K=4 was assumed, Deoni and Ongole clustered

    360 together independent of Kangayam ancestry while at K=5, Ongole clustered distinctly from

    361 that of Deoni. At K=6, Deoni and Bargur formed a separate cluster each, while Punganur and

    362 Vechur clustered together. At K=7, the miniature breeds, Pungaur and Vechur cattle clustered

    363 distinctly. Significant genetic admixture of zebu ancestry was observed among Alambadi,

    364 Pulikulam and Umblachery cattle consistently under the assumption of K=5 to 8.

    365 To further understand the degree of phylogeographic structure in the studied cattle breeds,

    366 analysis of molecular variance (AMOVA) was performed to assess the distribution of

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 15

    367 microsatellite variation as a function of both breed membership and geographic origin (Table

    368 3). When no grouping was assumed, 94.56% of the total variance among zebu cattle was found

    369 within breeds while 5.44% of total variance was explained by between breed differences. When

    370 the breeds were grouped according to their province of origin (Grouping I – based on

    371 state/province), about 0.32% of total variance was due to differences among groups (P>0.05)

    372 while 5.21% was due to differences among populations within groups. When the grouping was

    373 done based on the classification of zebu cattle by Joshi and Phillips (1953) (Grouping II),

    374 among group variance was negligible (0.28%; P>0.05), while most of the between population

    375 variance was within groups (5.26%; P0.05) while variance among populations within groups was 4.41%. When the grouping was

    378 based on Bayesian genetic structure (Grouping IV; Group1-DEO; Group2-ONG; Group3-

    379 BAR, ALA, HAL, PUL, UMB; Group4-VEC; Group5-PNR; Group6-KAN), among group

    380 variance was 2.78% and significant (P

  • 16

    392 informative sites. The overall nucleotide diversity and average number of nucleotide

    393 differences among South Indian cattle was 0.0111and 10.36 respectively. A total of 207 unique

    394 haplotypes were observed with overall haplotype diversity of 0.961 (Table 4). The haplotype

    395 diversity observed in most South Indian cattle breeds was relatively higher than that of North

    396 Indian zebu cattle located along the Gangetic plains (Sharma et al. 2015). Two major maternal

    397 haplogroups, I1 and I2, typical of zebu cattle were observed, with the former being predominant

    398 occurring 313 times and the later occurring 166 times. However, there were differences in

    399 certain breeds like Vechur and Umblachery in which I2 haplogroup was more frequent than I1.

    400 Interestingly, six zebu cattle showed taurine lineage, of which one belonged to T1 (Deoni),

    401 three belonged to T2 (one each from Hallikar, Bargur and Kangayam) and two belonged to T3

    402 (one each from Deoni and Pulikulam). The overall nucleotide diversity within I1 and I2

    403 haplogroups were 0.0024 and 0.0029 respectively while the overall average number of

    404 nucleotide differences were 2.32 and 2.77 respectively (Table 5). Among the 207 unique

    405 haplotypes, 105 belonged to I1 haplogroup, 74 belonged to I2 haplogroup, 3 belonged to T3

    406 haplogroup and one each belonged to T1 and T2 halpogroups (Table 4). The overall haplotype

    407 diversity within I1 haplogroup and I2 haplogroup was 0.907 and 0.949 respectively. The

    408 haplotype diversity within I2 haplogroup was higher than that of I1 in all the South Indian zebu

    409 cattle breeds except Alambadi. Among the observed haplotypes, 27 were found to be shared

    410 across 2 or more breeds, of which 18 belonged to I1 haplogroup, 8 belonged to I2 haplogroup

    411 and one belonged to T3 haplogroup (Supplementary Table ST4). The haplotype

    412 HxJxABDLKOPUVHJ_H13 was the most frequent among I1 haplogroup and shared with all

    413 the investigated populations except Punganur cattle. Among the I2 haplogroup,

    414 HxJxABDLKOPUV_H7 was the most frequent haplotype and shared among 11 populations

    415 with the exception of Punganur, Holstein-Friesian and Jersey cattle. A total of 139 singleton

    416 haplotypes were observed, of which 67, 50, 18, 3 and one belonging to I1, I2, T3, T2 and T1

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 17

    417 haplogroups respectively. The higher proportion of singleton haplotypes within I2 haplogroup

    418 of South Indian cattle was consistent with relatively higher haplotype and nucleotide diversity

    419 observed within this lineage.

    420 3.5. mtDNA phylogeny and evolutionary relationship

    421 Phylogenetic analysis was performed on all the 207 unique haplotypes along with reference

    422 sequences of each haplogroup, the results of which is presented in Supplementary Figure SF3.

    423 The phylogenetic tree conformed to haplogroup classification with the formation of two major

    424 indicine clusters I1 and I2. There were no major sub-clusters except for one minor clade

    425 containing four or more haplotypes in each of the I1 and I2 haplogroups. The minor clade

    426 within in I1 haplogroup was formed by four singleton haplotypes (K_H142, L_H120, L_H115

    427 and V_H181) from Kangayam, Hallikar and Vechur cattle breed. Similarly, the minor clade

    428 within I2 haplogroup was formed by three singleton and two shared haplotypes (P_H155,

    429 AD_H61, B_H84, L_H129 and OP_H149) from Pulikulam, Alambadi, Deoni, Bargur, Hallikar

    430 and Ongole cattle.

    431 The frequency of mtDNA control region haplotypes was further utilized to estimate pairwise

    432 genetic differentiation among zebu cattle from South India as well as the zebu inhabiting areas

    433 closer to Indus valley sites (West Indian zebu) and Gangetic plains (North Indian zebu and

    434 Nepalese zebu). Analysis of I1 haplotypes revealed pairwise FST varying between zero and

    435 o.293 while I2 haplotypes showed pairwise FST varying between zero and 0.255. The pairwise

    436 FST among I1 haplotypes was zero among NIC-NPC, WIC-NPC, ALA-UMB and HAL-UMB

    437 breed combinations and less than 1% among ALA-HAL, WIC-HAL, WIC-UMB, NIC-UMB,

    438 ALA-WIC breed combinations. The largest FST among I1 haplotypes was observed between

    439 Punganur cattle and other zebu breeds (Vechur, Ongole, Deoni, Bargur and Nepalese) followed

    440 by Vechur and other breeds (Pulikulam, Ongole and Umbalchery). In contrast, the pairwise FST

    441 among I2 haplotypes was zero between Punganur and other zebu breeds (ONG, BAR, NIC,

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 18

    442 UMB, DEO, HAL, WIC, ALA, KAN) and among HAL-DEO, WIC-DEO and PUL-DEO breed

    443 combinations. The largest FST among I2 haplotypes was observed between Nepalese zebu and

    444 South Indian zebu (ONG, KAN, PUL, BAR, UMB, VEC, DEO and ALA) followed by Vechur

    445 and other breeds (ALA, BAR, UMB, ONG, PUL, HAL and DEO).

    446 The pairwise FST for each of the haplogroups were further utilized to perform principal

    447 components analysis (PCA) and understand the evolutionary relationships among the zebu

    448 cattle breeds (Figure 4). Principal components analysis (PCA) performs orthogonal

    449 transformation of a set of correlated variables (genetic distance estimations) into a set of new

    450 linearly uncorrelated variables called as principal components. The three largest principal

    451 components that explained maximum variation in the dataset were used to project sampled

    452 breeds into a three-dimensional geometric space. PCA of I1 haplotypes revealed the three

    453 largest principal components explaining 93.04% of the total variance in the dataset. The three-

    454 dimensional scattergram drawn using the three largest principal components showed the

    455 distinctness of Punganur, Pulikulam and Vechur cattle from all the other South Indian zebu

    456 cattle breeds. The zebu cattle from North India and Nepal were also found to be distinct from

    457 West and South Indian zebu cattle. Similarly, PCA of I2 haplotypes revealed the three largest

    458 principal components together contributing to 91.31% of total variance in the dataset. The

    459 three-dimensional scattergram showed Vechur and Nepalese cattle distinct from other zebu

    460 cattle breeds.

    461 To understand the underlying genetic structure of each of the two maternal lineages (I1 and

    462 I2), analysis of molecular variance among zebu mtDNA haplotypes of cattle from South India,

    463 Gangetic plains (North India and Nepalese) and Indus valley (West India) sites were conducted

    464 (Table 6). When no grouping was assumed among I1 haplotypes, 6.17% of variance (P

  • 19

    467 India, Gangetic Plains and Indus valley sites), among group variance was negligible (0.43%)

    468 and insignificant (P>0.05). However, when Punganur, Vechur and Pulikulam breeds were

    469 grouped separately from other South Indian zebu breeds, among group variance was 2.91%

    470 and statistically significant (P0.05). However, when Vechur and Nepalese cattle were grouped

    477 together, among group variance increased to 6.04% and was statistically significant (P

  • 20

    492 most of the South Indian breeds with the modal value at two pairwise mismatches. Bimodal

    493 distribution was observed among I1 haplotypes in Kangayam, Umbalchery and Punganur

    494 breeds with modal values at 1 and 3, 1and 5, 2 and 6 mismatches respectively. In case of zebu

    495 cattle from Indus valley site (WIC), the modal value was observed at 3 mismatches while the

    496 Gangetic plains Zebu cattle, NIC and NPC showed modal values at 3 and 2 mismatches

    497 respectively. The modal values of South Indian I1 lineage observed in the present study is

    498 contrary to the reports made earlier (Chen et al. 2010). Similarly, with respect to I2 lineage,

    499 modal values were observed around two to four mismatches in most of South Indian and

    500 Gangetic Plain Zebu cattle (NIC and NPC) while it was observed at one mismatch in Indus

    501 valley site zebu cattle (WIC). Unimodal distribution of I2 haplotypes was observed in most

    502 South Indian zebu breeds except Deoni, Kangayam and Pulikulam. The mismatch analysis of

    503 pairwise differences revealed the parameter ‘tau’ (the coalescent time of expansion) ranging

    504 from 1.268 (Ongole) to 8.406 (Punganur) among I1 haplotypes while it ranged from 1.344

    505 (Pulikulam) to 4.182 (Kangayam) among I2 haplotypes of South Indian zebu cattle. The

    506 estimated sum of squared deviations (SSD) and the raggedness index was significant (P

  • 21

    517 of smaller magnitude was observed within the I1 haplogroup indicating a possible sub-lineage

    518 of I1. The South Indian zebu cattle were further subjected to tests for selective neutrality, the

    519 Tajima’s D and Fu’s FS. The Tajima’s D statistics was based on frequency spectrum of

    520 mutations while Fu’s FS was based on haplotype distribution. The results revealed negative

    521 Tajima’s D in all the investigated South Indian breeds for haplogroup I1, of which Deoni,

    522 Hallikar and Pulikulam were statistically significant (P

  • 22

    542 Alambadi, Ongole, Pulikulam) and medium to large (Deoni, Kangayam) size animals; draught

    543 characteristics varying from heavy load pulling capability (Hallikar, Umblachery) to speed and

    544 endurance in trotting (Kangayam, Bargur); utilities ranging from draught to penning for dung

    545 (Pulikulam, Bargur). Joshi and Phillips (1953) classified Zebu cattle of India and Pakistan into

    546 six major groups based on their morphology, breed history and phenotypic characteristics. The

    547 South Indian zebu cattle investigated in the present study were classified under three major

    548 categories: Group II, Short horned, white or light grey cattle with coffin shaped skulls

    549 (Ongole); Group III, Ponderously, built Red and White cattle (Deoni); and Group IV, Mysore

    550 type, compact sized cattle (Hallikar, Kangayam, Umblachery, Bargur, Pulilkulam and

    551 Alambadi). The bantam weight dwarf type cattle breeds, Vechur and Punganur did not fall into

    552 any of these three categories.

    553 4.1. Population status, effective population size and inbreeding estimates

    554 With rapid mechanization of agriculture in South India, the need for rearing draught cattle came

    555 down drastically in the last few decades. This led to reduction in effective population size and

    556 consequent increase in rate of inbreeding in many of these breeds (Singh and Sharma, 2017).

    557 A significantly strong negative Pearson’s correlation coefficient of -0.674 (P

  • 23

    567 had higher mean FIS estimates of 0.089 and 0.076 respectively. Although the total population

    568 of both these breeds were relatively higher (39050 and 14154 heads of Umblachery and Bargur

    569 respectively), only few hundred breedable males (586 and 624 for Umblachery and Bargur

    570 respectively) were present in their respective native breed tracts. Lack of availability of

    571 sufficient breeding bulls was one of the main reasons for low effective population size, higher

    572 estimates of FIS and high rate of inbreeding in these breeds (Singh and Sharma, 2017). Among

    573 the South Indian zebu cattle, the lowest FIS estimate was observed in Hallikar cattle

    574 (FIS=0.021), whose total population was greater than 1.2 million heads with breedable male

    575 and female population of 5524 and 397357 respectively. Interestingly, Punganur cattle with a

    576 total population of less than 3000 and effective population size of ~200 showed positive, but

    577 relatively lower estimate of FIS (0.041). This could be possibly due to a better breedable male

    578 to breedable female ratio observed in this breed (0.0715) as compared to Vechur (0.0020),

    579 Pulikulam (0.0084) and Umbalchery (0.0387) cattle (DADF, 2015).

    580 Singh and Sharma (2017) assessed the degree of endangerment of Indian zebu cattle breeds

    581 based on estimated breed-wise cattle population data (DADF, 2015), effective population size

    582 and estimated rate of inbreeding. Using the criteria laid down by Food and Agriculture

    583 Organization of the United Nations (FAO, 2013) and the Indian Council of Agricultural

    584 Research-National Bureau of Animal Genetic Resources (ICAR-NBAGR, 2016), they

    585 classified the status of Vechur as “critical” (breeding males less than five, breeding females

    586 494 and high rate of inbreeding), Punganur and Pulikulam as “endangered” (breeding females

    587 more than 300 but less than 3000) and Bargur as “vulnerable” (breeding females more than

    588 3000 but less than 6000). Thus, the status of four out of ten South Indian zebu cattle breeds

    589 investigated in the present study are in the category of “under risk” from conservation

    590 standpoint. The level of inbreeding estimates is consistent with their population status and is a

    591 cause for concern from conservation standpoint. It is also worth to mention that the population

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 24

    592 status of “Alambadi” breed is currently not known but has been observed to retain moderate

    593 levels of heterozygosity.

    594 4.2. Origin, breed history and population structure

    595 Most South Indian cattle breeds belong to Mysore type cattle (Group IV), of which Hallikar is

    596 considered to be the ancient one and widely distributed with significant and stable population

    597 size. Ware (1942) and Joshi and Phillips (1953) mentioned the origin of most Mysore type

    598 cattle breeds centred around Hallikar cattle. The results of AMOVA and genetic structure

    599 analysis obtained in the present study was in conformity with this theory of origin for most, if

    600 not all the breeds. Among group variance was significant (P

  • 25

    617 colour, typical of being red and white or red with white spots. These cattle are known for their

    618 speed and endurance in trotting, but are very fiery, restive and difficult to train (Ware, 1942;

    619 Ganapathi et al. 2009; Pundir et al 2009). It also needs to be noted that among these four Mysore

    620 type Tamil Nadu breeds (Alambadi, Bargur Umblachery and Pulikulam), significant selection

    621 process had gone into the development of Bargur cattle, particularly for their unique

    622 morphological characteristics, herd uniformity and adaptation to zero input, hilly environment.

    623 This was evidenced by their individual assignment to unique inferred cluster when K=7 to K=8

    624 (Figure 3) was assumed under Bayesian clustering analysis.

    625 Pulikulam, is a migratory cattle breed, maintained in a zero-input system in southern parts of

    626 Tamil Nadu state of South India. These cattle have predominantly white or greyish coat colour

    627 and are mostly used for agricultural operations and rural transport. These cattle are well-known

    628 for their utilities as a sport animal for Jallikattu (a traditional bull embracing sport) and penning

    629 (a practice of manuring by keeping the cattle overnight in open agricultural fields). These cattle

    630 resemble Kangayam in physical appearance but are significantly smaller and compact in size.

    631 Gunn (1909) opined that these cattle probably had a strain of Mysore cattle blood in them. The

    632 present study revealed Pulikulam as a genetically distinct breed from Kangayam but sharing

    633 significant ancestry with Hallikar and Deoni cattle (Figure 3; K=4 to K=7). The genetic

    634 difference between Pulikulam and Kangayam was also consistent with the reported differences

    635 in morphometric parameters that included body length, height at withers and chest girth

    636 (Kanakraj and Kathiresan, 2006; Singh et al. 2012).

    637 Umblachery is an excellent draught cattle breed reared in the eastern coastal areas of Tamil

    638 Nadu state (Rajendran et al. 2008). These are medium sized cattle, selectively bred for short

    639 stature and suitability to work in marshy paddy fields. Anecdotal evidence attributed the

    640 possible origin and development of Umblachery from Kangayam cattle. Gunn (1909)

    641 mentioned that with the exception of appearance of head of these animals due to dehorning,

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 26

    642 these cattle retained the chief characteristics of Kangayam. However, the present study showed

    643 significant genetic differentiation among these breeds (FST=0.070) and little or no traces of

    644 Kangayam ancestry in Umblachery cattle. Similar to Pulikulam cattle, significant admixture

    645 was observed in Umblachery cattle with traces of Hallikar and Deoni ancestry in them. This

    646 could be possibly due to trade influx of cattle into native tract of Umbalchery more than a

    647 century ago, from Southern Tamil Nadu (Cumbum district) where the now extinct breed called

    648 “Kappiliyan” or “Tambiran Madu” of Canarese (Mysore) origin was in abundance (Gunn,

    649 1909; Littlewood, 1936).

    650 The genetic structure analysis clearly indicated the influence of Hallikar breed in the

    651 development of most Mysore type cattle breeds of Tamil Nadu (Alambadi, Bargur,

    652 Umblachery and Pulikulam) with the exception of Kangayam. Kangayam is one of the popular

    653 draught breeds of South India known for their power (0.8hp per pair of bullocks) and sturdiness

    654 (Surendrakumar, 1988). Although, there are reports indicating the existence of Kangayam

    655 cattle for nearly three centuries (Reddi, 1957), the origin and development of this breed has

    656 been attributed mainly to 33rd Pattagar (community chief) of Palayakottai who selectively bred

    657 the local cattle, only with sires available in his herd. The appearance of some of his cows/heifers

    658 gave the impression that they had a distinct strain of Ongole blood in them, as the Pattagar was

    659 known to possess purebred Ongole cows (Gunn, 1909). But no records were maintained by

    660 Pattagar during the early stages of the development of this breed to confirm this. Kangayam

    661 cattle conform largely to Mysore type breeds, but with a relatively larger body size, probably

    662 indicating the admixture with a heavier breed (Phillips, 1944). However, the results of

    663 phylogeny, multidimensional scaling and analysis of molecular variance in the present study

    664 clearly showed the genetic uniqueness of Kangayam cattle. The pairwise FST indicated marked

    665 differentiation of Kangayam from Hallikar (FST=0.093) and Ongole (FST=0.075) cattle. Further,

    666 Kangayam was also significantly differentiated from other Mysore type breeds (Alambadi,

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 27

    667 Bargur, Umblachery and Pulikulam) with pairwise FST ranging from 6.3% to 7.7%. The

    668 Bayesian clustering analysis showed Kangayam as a genetically distinct entity from both

    669 Ongole and Hallikar cattle. Low levels of admixture observed in Kangayam cattle, when

    670 assumed under K=4 to K=8 (Figure 3), was probably due to ancestral zebu inheritance common

    671 to all South Indian breeds. The genetic uniqueness of Kangayam cattle might have been the

    672 outcome of systematic selective breeding practices of Pattagars initially, and subsequent

    673 introduction of herd book registration followed by implementation of various genetic

    674 improvement programs since1940s by Government of Madras (Panneerselvam and

    675 Kandasamy, 2008).

    676 The two bantam cattle breeds, Vechur and Punganur are unique to South India and are probably

    677 among the smallest cattle breeds of the world, with the average height of adults ranging from

    678 60-100 cm and body weight of 125-200 kg (Nath, 1993; Iype, 1996). Vechur and Punganur

    679 cattle have retained moderate levels of diversity despite the small population size and being

    680 classified in the critically endangered category. Both these breeds were distinct from rest of the

    681 South Indian breeds, not only in terms of morphology but also in terms of milk production

    682 efficiency. The cows are considered to be efficient milk producers relative to their size, with

    683 peak milk yield scaling up to 3-4 kg/day (Nath, 1993; Iype, 1996). Genetic distance based

    684 phylogenetic analysis and multidimensional scaling plot derived from pairwise FST showed

    685 Vechur and Punganur to be genetically distinct from other South Indian draught type zebu

    686 cattle. Although, Bayesian clustering analysis indicated the influence of Hallikar ancestry in

    687 both the breeds (K=3 to K=6), they clustered independently at K=7 and K=8 (Figure 3). It is

    688 also worth to mention that a high degree of genetic differentiation was observed among these

    689 two dwarf cattle (Vechur and Punganur) breeds (FST=0.092). This is understandable

    690 considering the geographical location of their respective native tracts. The isolation of native

    691 tract of Vechur cattle (Kottayam district of Kerala state) by rivers, canals and backwaters and

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 28

    692 the distant location of Punganur cattle habitat (Chittoor district of Andhra Pradesh state)

    693 coupled with selective breeding practiced by farmers, might have contributed to the genetic

    694 differentiation of the two breeds. In addition to restricted gene flow, genetic drift arising out of

    695 small population size of both the breeds could have also contributed to their genetic divergence.

    696 4.3. Breed purity and level of taurine introgression

    697 With the mechanization of agricultural operations and increasing demand for dairy products in

    698 the last few decades, the focus of cattle breeding in South India shifted towards improving milk

    699 production. Beginning with the Operation Flood program in 1970s, intensive cattle

    700 development projects in Southern India were targeted towards genetic improvement of local

    701 cattle for milk through crossbreeding with exotic commercial taurine cattle. Although the

    702 official breeding policy emphasized the need for selective breeding of recognized zebu cattle

    703 breeds, indiscriminate crossbreeding took place in the native tracts of many indigenous breeds.

    704 In the absence of pedigree recording under small holder production set up, it is not uncommon

    705 that the local cows inseminated with purebred taurine or crossbred semen were subsequently

    706 served by locally available sires or vice versa. Often, the choice of semen used for artificial

    707 insemination (AI) is left to the inseminators and depends on the availability of purebred zebu,

    708 purebred taurine or crossbred semen straws, irrespective of the breed/genotype of the cows

    709 (purebred zebu, non-descript local, crossbred cattle, etc.) brought for AI. This resulted in

    710 dilution of purebred zebu germplasm with varying levels of taurine admixture in well described

    711 zebu cattle populations, thus affecting their breed purity.

    712 Bayesian clustering analysis was performed to assess the taurine admixture in South Indian

    713 zebu cattle using purebred Jersey and Holstein-Friesian as reference genotypes. Samples from

    714 Bos taurus X Bos indicus crossbreds were also genotyped to evaluate the range of taurine

    715 inheritance under field conditions. The proportion of membership coefficient obtained for

    716 individual animals in the inferred zebu and taurine clusters were utilized to estimate taurine

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 29

    717 admixture and purity of the investigated zebu cattle breeds (Table 8). Ongole had zero

    718 individuals with >6.25% taurine admixture while Punganur had the highest proportion (44.4%)

    719 of individuals with >6.25% taurine admixture. Among the South Indian cattle, Ongole,

    720 Kangayam, Deoni, Alambadi and Bargur breeds showed majority of their individuals (>95%

    721 of each of these populations) with less than 6.25% taurine admixture. The relatively low level

    722 of taurine admixture in Ongole, Kangayam and Deoni can be attributed to the following factors:

    723 (i) better artificial insemination coverage in their respective breed tracts (ii) availability of

    724 purebred zebu, true to the type breeding bulls (iii) availability of purebred zebu semen for AI

    725 and (iv) better implementation of breeding policy in favour of selective breeding of purebred

    726 zebu cattle (Panneerselvam and Kandasamy, 2008; Natarajan et al. 2012; Dongre et al. 2017).

    727 In case of Alambadi and Bargur cattle raised in a unique, extensive production system of hilly

    728 areas, the access to artificial insemination services is little or absent. The cows in heat are

    729 normally served by the bulls that accompany the herd while grazing in the forest areas. For

    730 example, in Bargur cattle, the herd composition includes at least 2% of total size as breeding

    731 bulls and young male stocks (Ganapathi et al. 2009) in addition to 18-20% bullocks. The

    732 farmers select the breeding bulls based on certain phenotypic characteristics including

    733 uniformity of coat colour (Pundir et al 2009). The absence of AI facility and natural service

    734 being the main breeding practice in the native tract, the introduction of exotic bull semen was

    735 limited in Bargur and Alambadi cattle resulting in low levels of taurine admixture in these

    736 breeds. With respect to Hallikar cattle, the proportion of individuals with >6.25% taurine

    737 admixture was estimated to be 16.7%. This was surprisingly higher, considering the significant

    738 and stable population size of this breed (DADF, 2015). Singh et al. (2008) reported the demand

    739 for purebred Hallikar semen was higher than the available supply in the native tract and thus

    740 the lack of quality breeding bulls might have contributed to increased levels of taurine

    741 admixture in this breed.

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 30

    742 Breeds with highest proportion of individuals with >6.25% taurine admixture included

    743 Punganur (44.4%), Vechur (42.3%), Umblachery (33.3%) and Pulikulam (23.5%) cattle.

    744 Significant proportion of individuals belonging to these breeds also had >12.5% taurine

    745 admixture. Among these, Vechur and Pulikulam have been reported to have severe shortage of

    746 breedable males (DADF, 2015; Singh and Sharma, 2017) in their respective native tracts.

    747 However, the AI coverage in these areas is high and the semen of purebred, commercial taurine

    748 cattle are readily available for insemination in local cows. This might have resulted in

    749 indiscriminate crossbreeding of purebred zebu cattle, thereby increasing the level of taurine

    750 admixture in them. In case of Punganur cattle, although reasonable number of breedable males

    751 are available in the native tract, the level of taurine admixture is relatively higher, possibly due

    752 to well established AI programs and readily available exotic taurine semen for insemination.

    753 Kannadhasan et al. (2018) utilized community based participatory approach to identify and

    754 prioritize the constraints in Umblachery cattle farming: (i) shift in farmers’ preference towards

    755 crossbreds for better milk production (ii) lack of interest among farmers in using Umblachery

    756 bullocks for agricultural operations (iii) lack of grazing lands (iv) lack of quality purebred

    757 breeding bulls and (iv) inadequate efforts of breeders’ association in promoting and improving

    758 Umblachery breed. The above factors coupled with high coverage of AI and easy access to

    759 taurine/crossbred semen might have contributed to relatively higher levels of taurine admixture

    760 in Umblachery cattle.

    761 4.4. Maternal lineage diversity and zebu cattle domestication

    762 The present study revealed no additional maternal haplogroups in the South Indian cattle other

    763 than the typical I1 and I2 zebu lineages reported earlier (Chen et al. 2010). A relatively high

    764 average number of nucleotide differences, nucleotide and haplotype diversity was observed in

    765 I2 lineage as compared to that of I1lineage. Mismatch distribution analysis of mtDNA control

    766 region sequences revealed the modal value at two mismatches for I1 lineage, while it was

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 31

    767 observed at two to four mismatches for I2 lineage in most of the South Indian breeds. The

    768 mismatch distribution analysis thus clearly showed high levels of mitochondrial control region

    769 diversity in South Indian zebu cattle. In general, extant populations located nearby centres of

    770 domestication retain high levels of diversity, while populations located farther away from the

    771 domestication centre show reduced diversity. The high levels of mtDNA control region

    772 diversity observed in South Indian zebu cattle is interesting. Particularly, the pairwise

    773 differences among the I2 haplotypes of South Indian cattle were relatively higher than West

    774 Indian (Indus valley site) zebu cattle and comparable to that of North Indian and Nepalese zebu

    775 cattle (Gangetic plains zebu) (Chen et al. 2010).

    776 Several studies have demonstrated independent domestication and origin of Bos tauus and Bos

    777 indicus cattle based on mtDNA variations (Loftus et al. 1994). It is also clearly established that

    778 the humpless Bos taurus cattle were domesticated in the Near East during the Neolithic

    779 transition about 10 thousand years before present (Troy et al. 2001; Beja-Pereira et al. 2006;

    780 Edwards et al. 2007; Kantanen et al. 2009; Bonfiglio et al. 2010). Similarly, the report of Chen

    781 et al. (2010) shed much light on the origin and domestication of humped zebu cattle in the

    782 Indian sub-continent. They analyzed mitochondrial control region sequences from six different

    783 regions of Asia (i) Indus valley region (ii) Gangetic plains (iii) South India (iv) North East

    784 India (v) West Asia (vi) East Asia (vii) Central Asia and (viii) South East Asia. Among these,

    785 nucleotide diversity was reported to be notably higher in four regions of the Indian sub-

    786 continent (Indus Valley, Gangetic Plains, South India and North East India), indicating a vast

    787 expanse of area as the probable centre of domestication. Chen et al. (2010) identified the greater

    788 Indus valley (West India and present-day Pakistan) as most likely location for the origin and

    789 domestication of I1 maternal lineage. They made the above conclusion based on the following

    790 evidences: (i) the backbone structure of I1 haplogroup network being formed by cattle

    791 populations in the Indus valley region (ii) modal value of mismatch distribution of pairwise

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 32

    792 differences being higher in Indus valley cattle (two mismatches) as compared to those located

    793 in Gangetic plains (one mismatch) and South India (one mismatch) and (iii) relatively high I1

    794 haplotype diversity observed in Indus valley cattle as compared to cattle in Gangetic plains and

    795 South India. However, analysis of I2 haplogroup sequences showed more complex pattern of

    796 diversity among cattle located in these regions. Identical modal values of I2 mismatch and

    797 nearly identical haplotype diversity were observed in Indus valley and South Indian cattle while

    798 the frequency of I2a (an intermediate I1-I2 haplotype) was relatively higher in cattle from

    799 Gangetic plains. Hence, Chen et al. (2010) reported the lack of genetic evidence to pin-point

    800 a single location/region for the origin of the I2 haplogroup. However, they opined the possible

    801 initial domestication of this lineage in the northern part of the Indian sub-continent based on

    802 the chronology of available archaeological evidence for domesticated zebu cattle and probable

    803 wild aurochs in the Indian subcontinent.

    804 The present study based on comprehensive sampling of diverse breeds of South Indian zebu

    805 cattle, showed the mismatch modal value of I1 haplogroup to be higher (two mismatch) than

    806 reported by Chen et al. (2010). More interestingly, the mismatch distribution analysis of I2

    807 haplogroup sequences revealed modal values around two to four mismatches in most South

    808 Indian breeds. This was higher than the modal value observed for West Indian (Indus valley)

    809 cattle, but comparable to that of North Indian and Nepalese (Gangetic plains) cattle. The

    810 intermediate I1-I2 haplotype sequences (I2a) were also observed in certain South Indian cattle

    811 breeds like Pulikulam, Deoni and Kangayam. Further, it is worth to mention that the haplotype

    812 diversity within South Indian I2 lineage was higher in most South Indian cattle breeds.

    813 Principal components analysis of pairwise FST derived from frequency of I2 haplotypes showed

    814 the distinctness of Vechur, Kangayam, Punganur and Alambadi cattle, thus indicating the

    815 diversity of this lineage within South Indian cattle. The results showed the need for additional

    816 sampling of zebu breeds from North, East and North-East Indian cattle to explore and gather

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 33

    817 new evidence on the potential location for the origin and domestication of I2 lineage in the

    818 Indian sub-continent.

    819

    820 Conclusions

    821 The present study revealed moderate levels of genetic diversity existing in South Indian zebu

    822 cattle. It also established the genetic distinctness of Kangayam, Vechur and Punganur cattle

    823 from the rest of the South Indian zebu cattle. The influence of Hallikar breed was observed in

    824 the development of most Mysore type cattle breeds of South India with the exception of

    825 Kangayam. Analysis of mtDNA control region sequences revealed two major maternal

    826 haplogroups, I1 and I2, with the former being predominant than the later. The results indicated

    827 the need for additional sampling and comprehensive analysis of mtDNA control region

    828 variations to unravel the probable location of origin and domestication of I2 zebu lineage. The

    829 present study revealed two major concerns about the status, conservation and improvement of

    830 South Indian draught type zebu cattle breeds: (i) Four of the ten investigated zebu cattle breeds

    831 are under the risk of endangerment due to small effective population size and high rate of

    832 inbreeding (ii) the lack of sufficient purebred zebu bulls for breeding and increasing level of

    833 taurine admixture in the native breeds tracts. Recently, the state governments of South India

    834 have started establishing conservation units and research centres for improvement of breeds

    835 like Alambadi, Bargur, Kangayam and Umblachery. Parental stocks of these breeds are being

    836 established by procuring true to the type animals from small holder farmers. However, in the

    837 absence of pedigree records, it has been difficult to estimate genetic purity or taurine admixture

    838 in those animals. Considering the relatively higher levels of taurine admixture in some of the

    839 breeds, it is advisable to perform molecular/genomic evaluation of local cattle to improve breed

    840 purity in these units/centres. It also needs to be mentioned that additional efforts are necessary

    841 to make the farmers aware of scientific breeding practices and limit the levels of inbreeding

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 34

    842 and taurine admixture in purebred zebu populations. Availability of purebred semen for

    843 artificial insemination, incorporation of genomic/molecular information to identify purebred

    844 animals and increased awareness among farmers will help to maintain breed purity, conserve

    845 and improve these important draught cattle germplasms of South India

    846

    847 Acknowledgement

    848 The present study is part of the Coordinated Research Project D3.10.28 of Joint FAO/IAEA

    849 Division of Nuclear Techniques in Food and Agriculture, International Atomic Energy Agency

    850 (IAEA), Vienna, Austria. The funding provided by IAEA for internship training of the first

    851 author at Animal Production and Health Laboratory, Seibersdorf, Austria is gratefully

    852 acknowledged. The authors also thank Dean, Veterinary College and Research Institute,

    853 Namakkal, Tamil Nadu, India for providing necessary facilities for the study.

    854

    855 References

    856 Antao T, Lopes A, Lopes RJ, Beja-Pereira A, Luikart G. LOSITAN: a workbench to detect

    857 molecular adaptation based on a FST outlier method. BMC Bioinformatics. 2008; 9: 323.

    858 Bandelt HJ, Forster P, Rohl A. Median-Joining networks for inferring intraspecific

    859 phylogenies. Molecular Biology and Evolution. 1999; 16: 37-48.

    860 Beja-Pereira A, Caramelli D, Lalueeza-Fox C, Vernesi C, Ferrand N, Casoli A, et al. The origin

    861 of European cattle: evidence from modern and ancient DNA. Proceedings of National

    862 Academy of Sciences, USA. 2006; 103: 8113-8118.

    863 Bonfiglio S, Achilli A, Olivieri A, Negrini R, Colli L, Liott L., et al. The Enigmatic Origin of

    864 Bovine mtDNA Haplogroup R: Sporadic Interbreeding or an Independent Event of Bos

    865 primigenius Domestication in Italy? PLoS One. 2010; 5(12): e15760.

    866 doi:10.1371/journal.pone.0015760.

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 35

    867 Chen S, Lin B-Z, Baig M, Mtra B, Lopes RJ, Santos AM, et al. Zebu cattle are an exclusive

    868 legacy of the South Asia Neolithic. Molecular Biology and Evolution. 2010; 27 (1): 1-

    869 6.

    870 DADF. Estimated livestock population breed-wise based on breed survey 2013. Published by

    871 Department of Animal Husbandry, Dairying and Fisheries, Ministry of Agriculture,

    872 Government of India, New Delhi; 2015.

    873 de Oliveira JFC, Neves JP, Almeida EA, Steigleder CS, Moraes JCF, Gonçalves PBD, Weimer

    874 TA. Association between reproductive traits and four microsatellites in Brangus-Ibagé

    875 cattle. Genetics and Molecular Biology. 2005; 28 (1): 54-59.

    876 Dieringer D, Schlötterer C. MICROSATELLITE ANALYSER (MSA): a platform indpendent

    877 analysis tool for larger microsatellite data sets. Molecular Ecology Resources. 2003; 3

    878 (1): 167-169.

    879 Dongre VB, Gandhi RS, Salunke VM, Kokate LS, Durge SM, Khandait VN, Patil PV. Present

    880 status and future prospects of Deoni Cattle. Indian Journal of Animal Sciences. 2017;

    881 87 (7): 800–803.

    882 Edwards CJ, Bollongino R, Scheu A, Chamberlain A, Tresset A, Vigne J-, et al. Mitochondrial

    883 DNA analysis shows a near eastern Neolithic origin for domestic cattle and no

    884 indication of domestication of European aurochs. Proceedings of the Royal Society B.

    885 2007; 274:1377-1385.

    886 Evanno G, Regnaut S, Goudet J. Detecting the number of clusters of individuals using the

    887 software structure: a simulation study. Molecular Ecology. 2005; 14: 2611-2620.

    888 Excoffier L, Hofer T, Foll M. Detecting loci under selection in a hierarchically structured

    889 population. Heredity. 2009; 103: 285–298.

    890 Excoffier L, Laval G, Schneider S. Arlequin ver. 3.0: An integrated software package for

    891 population genetics data analysis. Evolutionary Bioinformatics Online. 2005; 1: 47-50.

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 36

    892 FAO. In vivo conservation of animal genetic resources. FAO Animal Production and Health

    893 Guidelines. No. 14. Rome; 2013.

    894 FAO. Molecular genetic characterization of animal genetic resources. FAO Animal Production

    895 and Health Guidelines. No. 9. Rome; 2011.

    896 FAO. The second report on the state of the World’s Animal Genetic Resources for Food and

    897 Agriculture, edited by Scherf BD and Pilling D. FAO, Rome; 2015.

    898 Felsenstein J. PHYLIP: Phylogeny inference package, version 3.5. Department of Genetics,

    899 Washington University, Seattle, Washington; 1993.

    900 Ganapathi P, Rajendran R, Subramanian A. Distribution and Population Status of Bargur cattle.

    901 The Indian Veterinary Journal. 2009; 86: 971–972.

    902 Grant WS. Problems and Cautions With Sequence Mismatch Analysis and Bayesian Skyline

    903 Plots to Infer Historical Demography, Journal of Heredity. 2015; 106 (4): 333–346.

    904 https://doi.org/10.1093/jhered/esv020.

    905 Gunn WD. Cattle of Southern India Vol III. Bulletin No. 60. Department of Agriculture,

    906 Madras, India; 1909.

    907 Hiendleder S, Lewalski H, Janke A. Complete mitochondrial genomes of Bos taurus and Bos

    908 indicus provide new insights into intraspecies variation, taxonomy and domestication.

    909 Cytogenetic and Genome Research. 2008; 120:150–156. doi: 10.1159/000118756.

    910 ICAR-NBAGR. Guidelines for Management of Animal Genetic Resources of India. National

    911 Bureau of Animal Genetic Resources (Indian Council of Agricultural Research), Karnal

    912 (Haryana), India, 163 p.; 2016.

    913 Iype S. The Vechur Cattle of Kerala. Animal Genetic Resources Information. 1996; 18: 59-63.

    914 Joshi NR, Phillips RW. Zebu cattle of India and Pakistan. FAO Agriculture Studies No. 19,

    915 Food and Agricultural Organization of the United Nations, Rome, Italy. pp. 14-29;

    916 1953.

    .CC-BY 4.0 International licenseperpetuity. It is made available under apreprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in

    The copyright holder for thisthis version posted January 21, 2021. ; https://doi.org/10.1101/2021.01.21.427560doi: bioRxiv preprint

    https://doi.org/10.1101/2021.01.21.427560http://creativecommons.org/licenses/by/4.0/

  • 37

    917 Kanakraj P, Kathiresan D. Study on the physical and productive characteristics of Pulikulam

    918 breed of cattle and Katchaikatty sheep. Final Report of the project submitted to SEVA,

    919 Madurai; 2006.

    920 Kannadhasan MS, Kathirchelvan M, Rajendran R. Identification and Prioritization of

    921 Constraints in Umblachery Breed Cattle Farming Through Participatory Approach.

    922 International Journal of Current Microbiology and Applied Sciences. 2018; 7(11):

    923 1100-1110. https://doi.org/10.20546/ijcmas.2018.711.128

    924 Kantanen J, Edwards CJ, Bradley G, Vinalass H, Thessler S, Ivanova Z, et al. Maternal and

    925 paternal genealogy of Eurasian taurine cattle (Bos taurus). Heredity. 2009; 103: 404-

    926 415.

    927 Kienzle, Ashburner JE, Sims BG. Mechanization for Rural Development: A Review of Patterns

    928 and Progress from around the World. In Integrated Crop Management (FAO) No. 20;

    929 Food and Agriculture Organization of the United Nations: Rome, Italy; 2013.

    930 Li M-H, Iso-Touru T, Laurén H, Kantanen J. A microsatellite-based analysis for the detection

    931 of selection on BTA1 and BTA20 in northern Eurasian cattle (Bos taurus) populations.

    932 Genetics Selection Evolution. 2010; 42:32. http://www.gsejournal.org/content/42/1/32.

    933 Littlewood RW. Livestock of Southern India. Government of Madras, Madras, India. pp. 51-

    934 69; 1936.

    935 Loftus R, MacHugh D, Bradley D, Sharp P, Cunningham P. Evidence for two independent

    936 domestication of cattle. Proceedings of National Academy of Sciences, USA. 1994; 91:

    937 2757-2761.

    938 Natarajan A, Chander M, Bharathy N. Relevance of draught cattle power and its future

    939 prospects in India: A review. Agricultural Reviews, 2016; 37 (1): 49-54.

    940 Natarajan A, Senthilvel K, Velusamy K. Productivity performance of Kangayam cattle. Indian

    941 Journal of Animal Sciences. 2012; 82 (11): 1440–1441.

    .CC-BY 4.0 International licenseperpetuity. It is made available under ap