Next Article in Journal
Flunixin Meglumine Enhanced Bone Fracture Healing in Rabbits Associated with Activation of Early Collagen Deposition and Enhancement of Vascular Endothelial Growth Factor Expression
Previous Article in Journal
Effect of Palpable Udder Defects on Milk Yield, Somatic Cell Count, and Milk Composition in Non-Dairy Ewes
Previous Article in Special Issue
The Conservation of Native Domestic Animal Breeds in Nordic Countries: From Genetic Resources to Cultural Heritage and Good Governance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Quest for Genes Involved in Adaptation to Climate Change in Ruminant Livestock

1
Department of Animal Science, Food and Nutrition—DIANA, Università Cattolica del Sacro Cuore, Via Emilia Parmense, 84, 29122 Piacenza, Italy
2
Department for Innovation in Biological, Agro-Food and Forest Systems–DIBAF, Università Della Tuscia, Via S. Camillo de Lellis snc, 01100 Viterbo, Italy
3
Research Center on Biodiversity and Ancient DNA—BioDNA, Università Cattolica del Sacro Cuore, Via Emilia Parmense, 84, 29122 Piacenza, Italy
4
Laboratory of Geographic Information Systems (LASIG), School of Architecture, Civil and Environmental Engineering (ENAC), Ecole Polytechnique Fédérale de Lausanne (EPFL), 1015 Lausanne, Switzerland
5
Impacts on Agriculture, Forests and Ecosystem Services (IAFES) Division, Fondazione Centro Euro-Mediterraneo Sui Cambiamenti Climatici (CMCC), Viale Trieste 127, 01100 Viterbo, Italy
6
Nutrigenomics and Proteomics Research Center—PRONUTRIGEN, Università Cattolica del Sacro Cuore, Via Emilia Parmense, 84, 29122 Piacenza, Italy
*
Author to whom correspondence should be addressed.
Equal contribution.
Animals 2021, 11(10), 2833; https://doi.org/10.3390/ani11102833
Submission received: 19 August 2021 / Revised: 21 September 2021 / Accepted: 23 September 2021 / Published: 28 September 2021
(This article belongs to the Special Issue Animal Genetic Resources Conservation: From Optimal to Feasible)

Abstract

:

Simple Summary

After domestication in specific regions, livestock followed human migrations and colonized the whole world. During this population expansion, human and natural selection, together with demographic events, molded the livestock genome leading to local breeds and populations able to produce milk, meat, wool and tractive power in many different agro-climatic conditions. The climate is changing, with temperatures and the frequency of extreme climatic events increasing, which affects livestock welfare and production efficiency, particularly of the highly productive breeds. Genomics is now able to explore the DNA of local breeds adapted to extreme environments in search of genes carrying signatures of selection for adaptation. This review summarizes methods used to accomplish this task, giving examples of results achieved and perspectives for future breeding.

Abstract

Livestock radiated out from domestication centres to most regions of the world, gradually adapting to diverse environments, from very hot to sub-zero temperatures and from wet and humid conditions to deserts. The climate is changing; generally global temperature is increasing, although there are also more extreme cold periods, storms, and higher solar radiation. These changes impact livestock welfare and productivity. This review describes advances in the methodology for studying livestock genomes and the impact of the environment on animal production, giving examples of discoveries made. Sequencing livestock genomes has facilitated genome-wide association studies to localize genes controlling many traits, and population genetics has identified genomic regions under selection or introgressed from one breed into another to improve production or facilitate adaptation. Landscape genomics, which combines global positioning and genomics, has identified genomic features that enable animals to adapt to local environments. Combining the advances in genomics and methods for predicting changes in climate is generating an explosion of data which calls for innovations in the way big data sets are treated. Artificial intelligence and machine learning are now being used to study the interactions between the genome and the environment to identify historic effects on the genome and to model future scenarios.

1. Introduction

Climate change is generally causing an increase in global temperatures (see Box 1). The most recent estimates [1] suggest that a 1.5 °C warming compared to the 1850–1900 baseline will be reached in the second half of the current decade, but, in addition, there are longer cold periods and increased levels of solar radiation [2,3,4,5,6] (Figure 1).
These changes affect both extensive and intensive farming systems [7,8]. The impact of environmental changes on animals affects their health, growth, and fertility as well as the diseases to which they are exposed. In addition, availability and types of feed may change because of the impact of the climate on the production and quality of grains, pasture and forage crops [9,10], which will affect nutrition as well as animal health and metabolism [11]. Livestock can adapt to gradual changes in environmental temperature. However, rapid changes or extended periods of extreme conditions reduce their welfare and productivity and are potentially life threatening. Therefore, the current rapid rise in global temperature is increasingly exposing livestock to stress in many countries. Some local breeds that have been kept in areas with adverse conditions, such as high temperature and humidity or drought, have become adapted over many generations; these breeds are an invaluable resource for research and breeding. It is urgent to understand the biological mechanisms underlying their adaptability, and, in particular, to identify genomic regions and genes that control such mechanisms in order to facilitate the rapid selection of livestock resilient to climate change. This review focuses on ruminants and on the current state of knowledge on genetics controlling adaptation.

2. Impacts of Climate Change on Livestock

With increasing global temperatures, more productive livestock are at greater risk (see Box 2), because they have higher feed intake and feed consumption, which is directly related to animal heat production [12]. Animals eat less to counteract high temperatures, and nutrients are prioritized to support maintenance rather than production and reproduction. In the central U.S., for example, severe losses of beef cattle kept in feedlots have been reported because of heat waves in summer and extreme snowstorms and wind in winter [13]. Climate related economic losses as a result of animal death and reduced performance have been seen [14]. Cattle, sheep, pigs and chickens reduce their feed intake by 3–5% for each unit increase in temperature above 30 °C [15]. Reproduction is particularly affected. Hahn [16] reported that conception rates in dairy cows are reduced by 4.6% per unit change above 70 in the temperature humidity index (THI) [17]. For beef cattle kept in range or pasture management systems, a decrease in pregnancy rates of 3.2% and 3.5% was observed for each unit increase in average THI above 70 and an increase in average temperature above 23.4 °C, respectively. Among environmental variables, temperature has the greatest effect on cow pregnancy rates [18].
Climate change further includes altered rainfall patterns that, combined with geographical factors such as soil type, affect crop production [19,20,21]. Drought reduces biomass [22], increases lignin accumulation in plant tissues, and reduces proteins, resulting in less digestible forages [23] and insufficient energy to meet livestock requirements [24,25]. Increased occurrence of prolonged drought is therefore of great concern to pasture-based livestock systems [23], especially those in environments which cannot support arable production [26].
Climate change influences the distribution of animal pathogen vectors and parasite range [27] which, together with the decreased immune response of animals under stress (triggered by cortisol), exposes livestock to higher risks of disease. Early springs, warmer winters and changes in rainfall distribution affect the seasons in which pathogens, parasites and vectors are present, potentially increasing proliferation and survival of these organisms. Bluetongue recently spread northward from Africa to Europe [28] as a consequence of climate-driven ecosystem changes and the associated expansion of the geographic range of the insect Culicoides imicola, the vector of the virus [29]. Other vectors such as the tick Rhipicephalus appendiculatus, which is the host for the protozoan pathogen Theileria parva, are predicted to shift their geographic range due to climate change, moving southward from central sub-Saharan Africa towards southern Africa [30]. Higher temperatures in Europe have increased parasite burdens such as helminths, with a shift from species traditionally found in temperate zones such as Ostertagia ostertagi to tropically adapted species, particularly Haemonchus contortus [31,32]. In addition to temperature, increased rainfall and humidity have affected the distribution of parasites. Leptospirosis in humans has been linked to transmission from livestock, with many outbreaks reported following extreme weather events around the world [33].
Box 1. Climate Data and Tools.
High resolution meteorological data are used to evaluate climate trends and variability and to predict the frequency of extreme events. Where meteorological data are not available, advanced climate modelling produces “Climate Reanalysis” datasets for a comprehensive description of the climate in three-dimensional grids. “Climate Reanalysis” has become an essential tool for modelling meteorological data to provide services to sectors dependent on climate assessments, forecasts and projections, including ecosystem management, agriculture, and livestock farming [34,35]. Climate modelling is also able to produce short- to long-term climate predictions (months to a few decades ahead), and projections extending over many decades at the global level.
Bioclimatic indicators allow the ever–increasing climate datasets to be combined and condensed and are valuable for both expert and non-expert users. Bioclimatic indicators from several global datasets are available from WorldClim [36], CHELSA [37], CliMond [38], ecoClimate [39], ENVIREM [40], MERRAclim [41], CMCC-BioclimInd [42] and the latest, KGClim [43]. The FAO (Food and Agriculture Organization of the United Nations), provides Global Agro-Ecological Zoning (GAEZ) indicators of the likely variation in agricultural resources over time. Agrometeorological indicators from 1979 to the present and agro-climatic indicators from 1951 to 2099 derived from Climate Reanalysis and projections are available from Climate Change Service (C3S) of the Copernicus programme [44]. Frequency, duration, timing and severity of extreme weather events can be calculated using indicators and indices for climate extremes such as those defined by the Expert Team on Climate Change Detection Monitoring and Indices (ETCCDI) [45,46].

3. Becoming Adapted

Archaeological evidence and molecular analysis of present-day DNA variation suggest that livestock were domesticated in specific regions of different continents. The Fertile Crescent region in Southwest Asia is one of these. Here the wild progenitors of cattle, sheep, goats and pigs progressively adapted to a closer relationship with humans and finally became dependent on human care. Archaeozoological and mitochondrial DNA diversity data have confirmed that domestication of these species occurred in a climatically homogeneous area around the Fertile Crescent, comprising South-eastern Anatolia and the Iranian Zagros Mountains [47,48,49]. After domestication, livestock followed human migrations and, with agricultural expansion, colonized the whole world [50,51]. Technological advances have facilitated the study of ancient DNA (aDNA) from well–preserved archaeological remains, which is shedding light on the spatiotemporal dynamics of domestication and on the physiological and neurobiological changes that livestock species underwent during the transition from the wild to a domestic existence, as well as on the subsequent adaptation to different environments and selection for functional traits [52]. For example, these studies, have shown that cattle and goat domestication took place over relatively large geographical areas and extended time frames [53,54,55], with frequent events of admixture and introgression, sometimes from several wild relative species [53,56]. Over millennia livestock species have adapted to thrive in a range of environments, with different temperature, humidity, water and fodder availability and quality, pathogen and parasite challenges, and also to satisfy human needs for food, wool, fibre and tractive power.
At the genetic level, signals of adaptive changes driven by domestication have been found in genes related to nervous system development [57,58] including kit ligand (KITLG), the treacle ribosome biogenesis factor 1 (TCOF1), and fibroblast growth factor receptor 1 (FGFR1) [57]. Other signatures of selection, or of adaptive introgression from wild relatives, have been found in genes implicated in adaptation to feed and farming regimes. A variant in the cytochrome P450 2C19 gene (CYP2C19) has been under positive selection in goats. CYP2C19 is a member of the CYP2C subfamily of the cytochrome P450 superfamily of genes [59] which confers protection against a mycotoxin produced by Fusarium spp. fungi in cereals [54]. Therefore, the increased frequency of the variant is most likely a response to an increasingly cereal-based diet contained in waste by-products. Alleles that may have been introgressed into domesticated goats from Capra caucasica [60], a West Caucasian tur–like species, have been found in a genomic region harbouring genes that affect immune function and parasite resistance, including SERPINB3, SERPINB4, CD1B, COL4A4, BPI, MAN2A1, and CD2AP. In particular, the mucin 6 oligomeric mucus/gel–forming gene (MUC6), which encodes a gastro-intestinally secreted mucin, is nearly fixed in goats for the Tur–derived haplotype, which confers enhanced immune resistance to gastrointestinal pathogens [56]. The fixation of this introgressed variant may be the consequence of the adaptive advantage it provided in farm environments, where there is increased exposure to parasites and disease [56].
Recently, the characterization of the paleo-epigenome and paleo-microbiomes of domestic species have facilitated the exploration of their role in the adaptation of mammalian livestock to their environment [61]. Data on the epigenomic profiles or microbiota composition in ancient livestock may provide information on diet, lifestyle, health status and exposure to stressors, and thus help us to explore the mechanisms of adaptation and interaction with the environment on a micro-evolutionary scale.
Animals adapt to the environments in which they live and to external stress by acclimation to a particular stressor or to a range of stressors [62,63]. Adaptation can be crucial for survival, but often negatively affects the productivity and profitability of livestock systems. The ability to adapt depends in part on the flexibility of behavioral traits [64] and in part on morphological and physiological changes that better adapt animals for survival. For example, about 25% of sheep in the world are fat tail or fat rump breeds that are adapted to harsh semi-arid desert conditions where food availability is sporadic. The fat tail or rump acts as a store, to enable the animals to survive long periods when food is in short supply [65].
Cattle adapted to prolonged heat stress have increased hemoglobin and red cell numbers [66], which may also protect them against blood borne parasites such as theileriosis. Bos taurus taurus cattle that have been raised over many generations in cool and temperate climates have long hair, subcutaneous fat, and often a dark coat colour. In contrast, Bos taurus indicus cattle that were originally from hotter tropical climates have short hair, little subcutaneous fat, low metabolism, and a body conformation to aid heat dispersion, with high surface to volume ratio, large ears and loose skin, especially around the dewlap [67,68]. To increase performance while maintaining environmental resilience, crosses between taurine and indicine cattle have been developed [69]. The crossbred animals show better adaptation to high temperature and humidity, and to parasites, e.g., resistance to Boophilus microplus ticks increases in proportion to Bos taurus indicus ancestry in the cross [70].
Box 2. Heat Stress Indicators.
The level of heat stress experienced by an animal is the result of a combination of air temperature, relative humidity [71] and other climate factors including wind speed and solar radiation [72]. Depending on the management system, these parameters may make different contributions to the risk of thermal stress [73]. Environmental parameters can be measured and used to construct indices and set thresholds to define risk situations.
Most of the indices defining thermal stress risk have been developed for cattle, especially for dairy cows that are particularly susceptible to high temperatures. The Temperature Humidity Index (THI) [17] takes into account the effect of air temperature and humidity. THI was originally developed as a general indicator of heat stress for humans, but today is also applied to livestock. Over the years, the model and threshold values used to define heat stress conditions have been modified [73], and corrections are now applied if cooling systems are used in the housing [74]. THI does not take into account the cumulative effect of high temperature [73] or the impact of wind speed and solar radiation, which are important when estimating the level of heat stress experienced by an animal. The Equivalent Temperature Index (ETI) includes air speed in the formula [75], although solar radiation is not considered [73]. The THI adjusted (THIadj) index considers both the wind speed and the solar radiation, as well as breed and coat colour [76]. The Respiration Rate index (RR) is an extension of THIadj that also takes into account whether animals are in a shaded area or under the sun [77].
Other prediction models that have been developed to overcome the limitations of THI include the heat load index (HLI), which incorporates “black globe” temperature measurements substituting air temperature, animal factors (genotype, coat colour and health status) and management strategies (shade availability, days on feed, manure management and temperature of drinking water). These factors are used to modify the threshold to define the heat stress, and combined with factors to account for location-specific variables in different geographic areas [78]. HLI is considered a better predictor than THI as it includes the interaction between climatic variables and animal thermal exchange mechanisms [78]. The Accumulate Heat Load Unit (AHLU) index, based on HLI, is a measure of the animal’s heat load balance [79]. The AHLU may increase or decrease over time depending on HLI values. A zero AHLU value indicates that the animal is in thermal balance [79]. The HLI has also been extended to create a Comprehensive Climate Index (CCI) that can also be used under cold conditions [80].
A comprehensive review of models for predicting heat stress response in livestock is given in Rashamol et al. [78].
Senepol cattle were developed on the island of St Croix to create a breed that was polled, easily managed and tolerant of the tropical environment by crossing red polled taurine cattle with African Zebu cattle [81]. Some of these cattle have very short hair and reduced follicle density, giving the phenotype referred to as “SLICK”. SLICK is controlled by a single genetic locus and carriers of the Slick variant have lower core temperature than non-SLICK contemporaries [82]. Interestingly, the effect of SLICK is most likely through increased sweat production rather than the decrease in hair length and density [83]. The SLICK variant in Senepol cattle was initially mapped to chromosome 20 [84], and later the causative variation was identified in the prolactin receptor gene (PRLR). A single base deletion in exon 10 causes a frameshift that introduces a stop codon and results in the truncation of the protein [85]. Other criollo cattle breeds, such as Carora and Limonero, that were brought to the Americas from Spain 500 years ago [86] display a similar SLICK phenotype. However, these breeds do not carry the same prolactin variant that was identified in the Senepol cattle, although a genome-wide association analysis located the causative variant in or near to PLRL. DNA sequencing of SLICK Limonero cattle revealed three variants within the prolactin receptor gene that create premature stop codons in exon 11, one of which is also found in SLICK Carora cattle [87]. Recently, three novel variants were discovered in the PLRL gene in six Caribbean Basin cattle breeds. All create premature stop codons and increase heat tolerance. The occurrence of mutations in the prolactin receptor in several cattle breeds that are adapted to tropical climates and that have distinct evolutionary histories is unlikely to be by chance. Indeed, prolactin levels have been shown to be involved in thermoregulation in humans [88], showing that certain physiological processes and specific genes can be targeted by environmental pressure. The SLICK variant has now been introgressed into other breeds, including the highly productive Holstein dairy breed, creating more heat tolerant animals [89].
Nevertheless, adaptation generally requires changes in the combination of alleles of many genes; for example, the genomic analysis of admixture between Bos taurus taurus and Zebu (Bos taurus indicus cattle) in Africa showed that more than 150 loci were under selection for local adaptation [90]. The ability of livestock to successfully adapt to extreme climatic conditions and to tolerate a wide range of parasites has resulted in local populations with specific characteristics. These populations are valuable resources that, if well characterized, could be exploited to create breeds suited to new conditions arising from climate change.
Box 3. The Genome and Genomics.
The publication of the human genome sequence in 2001 [91] was a landmark that opened new opportunities in molecular genetics. The same approach that was used to sequence the human genome was used to produce draft sequences for the major livestock species; the first was the chicken in 2004 [92], followed by the cow in 2009 [93], then the pig [94], sheep [95] and goat [96] in 2012. These genomes became references against which DNA and RNA sequences from these species were aligned and compared. With the rapidly advancing sequencing technologies, which progressed from automated Sanger sequencing to next-generation high throughput short read sequencing [97], large numbers of individuals were sequenced at low resolution. Alignment of these sequences with the reference genomes revealed huge numbers of variations among individuals, in particular, Single Nucleotide Polymorphisms (SNP). This SNP data led to the development of genome-wide genotyping panels. A range of low (few thousand) to high (many hundred thousand) density SNP panels is commercially available, including some targeted to specific traits, and others that include SNP for several species to reduce costs of genotyping. Knowledge of the genome sequence from large numbers of individuals in a population enables low density SNP genotype data to be used to estimate higher density genotypes by “imputation” [98].
The analysis of phenotype and genotype in genome-wide association studies enables genetic loci with a major effect on the phenotype to be identified (e.g., [99,100,101]). In some cases the genes and causative polymorphisms controlling variations in target traits have been identified (e.g., [102]). Perhaps the most important advance coming from the availability of genome-wide SNP panels is that the idea of genome-based selection envisioned by Meuwissen and colleagues more than a decade ago has now been realized [103]. Other applications of the SNP panels include the analysis of population structure, history and diversity (e.g., [104,105,106] to guide conservation strategies [107] and the identification of regions of the genome that are under selection (e.g., [108]).
Next generation sequencing (NGS) has also facilitated the study of gene expression by enabling the analysis of the whole transcriptome [109]. Depending on how samples are processed and analysed, this approach can examine the expression of genes (e.g., [110,111]), variations in splice sites [112], and non-coding RNAs [113,114] as well as short, micro-RNAs [115] that have a regulatory role.
Further advances in sequencing technology are opening new opportunities. Long read, single molecule sequencing has enabled haplotype resolved genome sequences to be produced by separating the sequence reads originating from the maternally and paternally inherited chromosome [116,117]. Long read technologies such as Pacific Biosciences and Oxford Nanopore can produce full length sequences of transcripts to reveal isoforms present in different tissues or diverse physiological states. These technologies are also able to distinguish modified bases in the DNA, specifically methylation, in order to examine epigenetic patterns directly and explore the regulation of gene expression [118]. The Functional Annotation of Animal Genomes Consortium [119] is assembling data on genome structure, expression, and regulation using a range of new technologies. For an extensive review of the state of livestock genomics see Georges et al. [120].

4. Seeking Adaptive Genes

Several molecular genetic approaches have been used to identify adaptation-related genes. Genome wide association studies (GWAS) use phenotypes related to adaptation recorded directly on the animals. Landscape Genomics approaches use environmental variables as proxies for phenotypes. Other methods analyse the patterns of genomic diversity within and between populations and the level of admixture in specific genomic regions to identify selection signatures of adaptation. These approaches use genomic tools that may focus on individual loci through to whole genomic sequence analyses (see Box 3) and dedicated software (Table 1).

4.1. Genome-Wide Association Studies

Genome-wide association studies (GWAS) identify the association between variations in the genome, the genotype, with variations in phenotype displayed by individual animals belonging to a same breed or population. GWAS therefore requires both genotype and phenotype data on each individual [121,122]. Fulfilling such conditions is difficult for complex phenotypes, and not always feasible when the target population is small or isolated [123], which is often the case in adaptation studies. Moreover, costs for genotyping and trait recording represents a further hurdle in reaching an adequate sample size. For these reasons, GWAS carried out in livestock to understand the genetic control of complex traits, are invariably low powered and results between studies on the same traits are often inconsistent. In addition, the genetic associations identified are likely to differ depending on the way that a trait is measured, the genetic background and the environment. Livestock GWAS have primarily been used to identify genetic variants associated with specific production traits or disease responses [124]. GWAS that identify the genes controlling climate adaptation traits (e.g., efficient thermoregulation, feed utilization, and immunity) would accelerate selection for animals more resilient to climatic challenges [125].
Several statistical tests have been applied to identify marker–trait associations in GWAS, from single marker regression, to mixed model and Bayesian approaches that use different marker effect distributions as prior information, to haplotype based GWAS [126]. In all cases, corrections have to be applied for multiple testing and for population structure in order to avoid a high number of false positives. As most traits involved in adaptation are highly complex and have a low to moderate heritability, a large cohort of animals has to be investigated to reach a sufficient statistical power in GWAS. [127,128].
A GWAS of cattle indigenous to Benin [99] identified several potential candidate genes associated with stress and immune response (PTAFR, PBMR1, ADAM, TS12), feed efficiency (MEGF11, SLC16A4, CCDC117), and conformation and growth (VEPH1, CNTNAP5, GYPC). The study of cold stress in Siberian cattle breeds identified two candidate genes (MSANTD4 and GRIA4) on chromosome 15, putatively involved in cold shock response and body thermoregulation [100]. GWAS in taurine, indicine and cross-bred cattle identified PLAG1 (BTA14), PLRL (BTA20) and MSRB3 (BTA5) as candidate genes for several traits important for adaptation to extensive tropical environments [101]. A GWAS of the Frizarta dairy sheep breed, which is adapted to a high relative humidity environment, identified 39 candidate genes associated with body size traits including TP53, BMPR1A, PIK3R5, RPL26, and PRKDC [129]. An association analysis of genotype-by-environment (GxE) interactions with growth traits in Simmental cattle showed that birth weight was affected by temperature, while altitude affected weaning and yearling weight. Genes implicated in these traits included neurotransmitters (GABRA4 and GABRB1), hypoxia-induced processes (PLA2G4B, PLA2G4E, GRIN2D, and GRIK2) and keratinization (KRT15, KRT31, KRT32, KRT33A, KRT34, and KRT3), all processes that play a role in physiological responses associated with adaptation to the environment [130].
Enhancing efficiency would reduce the impact of changes in feed availability on livestock systems and potentially reduce methane production, which contributes to climate change. Residual feed intake (RFI), that is, the difference between actual feed intake and the theoretical energy requirements of an animal [131], has been used to select for increased feed efficiency (FE) [132,133]. A GWAS of RFI in Nellore cattle identified QTL on chromosomes 8 and 21 affecting the trait. Putative candidate genes on BTA 8 are CCDC171 and CLCN3 [134], while candidates on BTA11 are DEPP1, expression of which is induced by fasting, TUBB3 and PTSG1 [135].
A GWAS for temperament scores carried out on crossbred steers in a feedlot identified five SNP on BTA 1, 24, and 29 and 13 SNP on BTA11 [136]. Functional candidate genes close to these loci had roles in neural function included synaptotagmin 4 (BTA 24), FAT atypical cadherin 3 (BTA 29), tubulin tyrosine ligase-like 1 (BTA 5), spermatogenesis associated 17 (BTA 16), stanniocalcin 2 (BTA 20), and GABAA receptor γ 3 (BTA 21). A GWAS of 3,274 Charolais beef cows detected four significant and 12 suggestive chromosomal regions associated with several functional and behavioral traits including aggressiveness [137]. A recent GWAS analysis of 1,370 Brahman cattle clustered in two groups of temperament identified nine SNP located in intergenic regions near candidate genes ACER3, VRK2, FANCL [138].

4.2. Selection Signatures

Natural or artificial selective pressure causes an increase or decrease in the frequency of genetic variants in a population. Selection can be positive, balancing, or negative [139]. Positive selection increases the frequency of fitness-enhancing variants in a population whereas negative selection removes unfavourable mutations to restore DNA functional integrity [140]. Balancing selection retains more than one allele of a gene where heterozygotes have higher fitness [141]. The genes in the genomic region in linkage disequilibrium with the genes under selection will also increase or decrease in frequency through the hitch-hiker effect [142], changing the expected patterns of molecular variation and giving a “selection signature”.
Tajima’s D statistic (See Box 4) has been used to analyse wild and domestic sheep data to identify a genomic region involved in the resistance to pneumonia [143]. A scan of Russian cattle genomes using Tajima’s D statistic detected signatures of selection most likely resulting from adaptation to cold environments [144]. Fay and Wu’s H statistic has been used with cattle data to detect signals of recent positive selection involving genes associated with innate immune response [145].
Signatures of recent selection associated with aggressiveness have been identified on chromosome X by comparing the Lidia cattle breed, which has been selected for aggressive responses, with two Spanish breeds showing docile behaviour. The most significant selection signature included the monoamine oxidase A gene (MAOA) [146]. A further refinement of the analysis identified a variable number of tandem repeats in the gene, with the Lidia breed having fewer repeats compared with the docile breeds [147]. Favourable genetic and phenotypic relationships between docility and meat quality, feedlot performance, ease of transport and reproductive traits have been reported [148]. Temperamental animals generally are not as well adapted to stress and have slow growth rates, poor carcass conformation and poor immune function [149,150]. Differences in docility have also been found between Bos taurus taurus and Bos taurus indicus cattle (e.g., [151] and between beef and dairy breeds [152].
Signatures of selection related to feed adaptation have been found in sheep using an FST approach [153]. Of the seventeen genes under climatic selection, nine were related to energy metabolism. The strongest selection signal was around TBC1D12, on OAR22, which plays a role in GTPase regulation. The FST approach was also applied to Siberian cattle populations in order to understand the genetic basis of adaptation to cold environments [154]. Results identified several genes that have been implicated in thermal adaptation in cattle, such as GRIA4, COX17, MAATS1, UPK1B, IFNGR1, DDX23, PPT1, THBS1, CCL5, ATF1, PLA1A, PRKAG1, and NR1I2.
With regard to hot environments, Li and colleagues [155] investigated selection signatures of bovine heat tolerance in Dehong cattle, a Chinese indigenous zebu breed, using an FST approach. Results indicated that genes involved in heat shock (HSF1), oxidative stress response (PLCB1, PLCB4), coat color (RAB31), feed intake (ATP8A1, SHC3) and reproduction (TP63, MAP3K13, PTPN4, PPP3CC, ADAMTSL1, SS18L1, OSBPL2, TOX, RREB1, and GRK2) may play a role in heat adaptation.
Pairwise comparison of genetic differentiation of sheep breeds adapted to different environments identified selection signatures in the genes MITF, FGF5, MTOR, TRHDE and TUBB3 that have been associated with high-altitude adaptation [156]. An FST statistic approach applied to cattle breeds reared in different environments identified several genes under positive selection for thermal tolerance [157]. HapFLK detected the Nebulin Related Anchoring Protein gene (NRAP) to be under selection for adaptation to cold environments [158], ACSS2, ALDOC, EPAS1, EGLN1 and NUCB2 to be under selection for high-altitude adaptation in cattle [159], and DNAJC28, GNRH1 and MREG to be associated with heat stress adaptation in sheep [160].
iHS methods have been used to detect signatures of adaptation to environments in French Charolais cattle, sheep and goats [161,162,163,164,165]. Cross-population EHH-based tests have been used to detect hot climate adaptation in cattle [166,167,168] and sheep [169,170,171], and hypoxia adaptation in new world camelids [172]. Detecting runs of homozygosity (ROHs) to find regions containing genes associated with adaptation has been demonstrated in several domestic species [157,162,163,173,174].
Box 4. Approaches for Selection Signature Detection.
Selection on a locus, whether artificial for production or natural for adaptation, is associated with the reduction of genetic diversity in the region, creating a “selection signature”. Tajima’s test [175] is able to detect positive selection sweeps that occurred recently, as it identifies regions with high numbers of rare, low-frequency variants that are the result of recent mutation [176]. Fay and Wu statistics [177], in contrast, assess the relationship between ancestral and derived alleles, which enables both positive and negative recent selection occurring in medium- to high-frequency alleles to be detected. However, knowledge of ancestral alleles is necessary to apply the method [178].
Various approaches have been used to assess positive and negative selection in populations. Wright’s fixation index (FST) measures differences in allele frequencies between populations based on individual loci. FST has been used in many studies of livestock to explore differences among populations. A more recent approach to analyse population differentiation is the hapFLK metric [179], which improves on single locus statistics by testing haplotype differentiation. hapFLK corrects frequency estimates, accounting for the genetic relationship between populations using Reynolds genetic distances.
Selection for a favourable allele of a gene increases the levels of linkage disequilibrium (LD) around the locus under selection, until recombination occurs to reduce the extent of LD [180]. Selection signatures can therefore be found by detecting regions of strong LD relative to their prevalence within a population [181,182]. Alleles at linked loci are referred to as haplotypes. Extended haplotype homozygosity (EHH) methods measure the decay of haplotype homozygosity as a function of genetic distance. The integrated Haplotype Score (iHS) [183] is calculated from the integrals of the observed decay of EHH for the ancestral and derived alleles surrounding the locus under selection. Divergence between values from the genomic average is indicative of selection. This approach requires phased data and knowledge of the ancestral state for each allele, and it has low power when one allele is at high frequency or fixed. Cross-population methods such as XP-EHH [182] and Rsb [184] calculate EHH profiles between two populations, removing the need to know the ancestral state. These methods have high power for detecting selective sweeps that have reached fixation. Selective sweeps generate runs of homozygosity (ROH) when both parents pass on the same haplotypes that are inherited from one generation to the next [185].

4.3. Local Ancestry Inference

Local ancestry inference (LAI) identifies the ancestors of each genomic region at the chromosome level. LAI is also described as local ancestry deconvolution or chromosome painting. Local ancestry information can help to understand fine scale admixture and the population genetic history, identify recent targets of selection, guide the selection of reference panels for genotype imputation, and improve the detection power of genetic association studies of admixed populations [184,186,187,188,189]. Identifying the ancestry of chromosomal segments in admixed individuals facilitates the accurate identification of the history of genetic variants under selection [188], particularly where adaptive introgression has fixed or nearly fixed regions of the genome with specific population ancestry [190].
Most approaches to profile local ancestry divide the genome into windows and assign ancestry to each window by comparing it against a reference panel [186,188,191,192,193,194,195]. New methods do not require the explicit definition of a reference population [196,197]. The most popular algorithms for LAI rely on hidden Markov models (HMM), an extension of a Markov chain, to identify the transformation of a genomic region from the reference, which is often not obvious [198]. These methods provide the posterior probabilities for each possible ancestry state at each ancestry-informative site along the chromosome [189,190]. The estimates obtained depend largely on reference populations; therefore, approaches to identify convergent signals of ancestry across multiple tests using different references have been developed [199].
LAI has been widely applied to identify adaptive introgression related to climatic stressors in livestock. Adaptive introgression from wild to domestic sheep of loci affecting climatic adaptation and resistance to pneumonia has been identified using LAI [143,199]. Using LAI and multiple-reference adjustments, ancestry components of indicine origin were found in cattle breeds from Central Italy that are associated with resilience to harsh environments and climatic conditions [200]. A region of indicine introgression into Italian local taurine breeds has been identified on BTA18 containing KLHL36, USP10, KIAA0513 and FAM92B, all of which are related with residual feed intake [200]. This introgression could provide an adaptive advantage enabling animals to use low quality feed efficiently.
Introgression of genes regulating the response to hypoxia from yak into Tibetan cattle that facilitated the adaptation of the latter to high altitude was also identified by LAI [201]. Similarly, adaptive introgression of genes related to oxygen transportation from Argali sheep to Tibetan domestic sheep may be a key factor conferring high-altitude resilience [202]. Local ancestry signals in African cattle have identified the genomic components of indicine cattle related to heat tolerance and water reabsorption, along with innate-immune resistance to tick and tick-borne diseases [203]. LAI tests have provided evidence of adaptive introgression between llama and alpaca for coat colour, fibre characteristics, and adaptation to high altitude and harsh environment [172].

4.4. Landscape Genomics

Landscape genomics explores the interaction between the genome and the environment to better understand evolution by combining landscape ecology and population genetics [204,205]. Two advances enabled landscape genomics to be realized. The first was the development of Geographic Information Systems (GIS) [206], which facilitated the overlay of diverse geo-referenced information, in this case genetic and environmental data. The second was the availability of large numbers of genetic markers, specifically single nucleotide polymorphisms, that are easily assayed. The development of the software MatSAM to compare a large number of allele frequencies with eco-climatic variables brought these two advances together as landscape genomics [207]. The MatSAM software [208] has been successfully used for landscape genomics analyses of plant and animal species, including sheep [207], goats [209] and fish [210]. These studies used GIS to store both genetic and environmental variables retrieved from open access databases to create gene–environment matrices that are processed by logistic regressions. Several software programs using different models have been developed for land-scape genomic analysis; improvements of these have an ever-increasing capability to efficiently analyse big data sets of genomic and environmental variables (see Box 5).
Landscape genomics approaches were used to understand the genetic adaptation of South African goats, finding that climatic variables explained 17% of their overall diversity. Using SAM software (see Box 5 and [207]), 843 SNPs were identified that were associated with longitude, while LFMM software [211] found that 714 SNPs were associated with temperature and precipitation [212], with only one locus in common that included DGKB. These SNPs were close to genes involved in 205 biological pathways, all of which are potentially related to adaptation. Among the genes identified, several have been associated with thermoregulation in hot environments (e.g., PLCB1). In the analysis of a goat database of more than 1000 animals covering 33 Italian populations using landscape genomics methods and LFMM [213], identified many loci putatively associated with environmental variables, although there was no overlap in loci identified by each of the methods. Samβada identified 62 genes associated with temperature or precipitation; among these, RYR3 has been associated with mean temperature and ANK3 and BTRC with longitude [214]. The LFMM analysis identified four SNPs associated with Mean Diurnal Range and Mean Temperature. These SNP were near NBEA, located within a region involved with wool production in sheep [215], and RHOBTB1, which is known to be associated with meat quality in cattle [216]. As observed before, methods implemented in Samβada and LFMM produce non-overlapping results. The two software are suited to the analysis of population having specific genetic structure (see Box 5) and their use is suggesed as complementary rather than alternative tools. Colli et al. [217] applied landscape genomics software based on the SAM approach to analyse 43 European and West Asian goat breeds. Using AFLP markers, four loci were identified that were significantly associated with diurnal temperature range, frequency of precipitation, relative humidity and solar radiation.
A landscape genomic analysis of 57 sheep breeds using the SAM approach found that the DYMS1 microsatellite locus was associated with the number of wet days, which largely affects parasite load [207]. In an earlier study this locus was shown to be associated with parasite resistance [218].
Box 5. Landscape Genomics Software.
With the availability of increasing numbers of measures of environmental variables and an increasing number of genetic markers, the MatSAM software [208] was developed to process many simultaneous univariate association models. Samβada [213] is able to compute univariate and multivariate logistic regressions, integrate and make an intelligent selection of significant models, calculate pseudo R2, Moran’s I, and Geographically Weighted Regressions. This software has High Performance Computing (HPC) capacities to handle the large datasets created when several million SNPs, produced by high-throughput sequencing, are combined with hundreds of environmental variables. Samβada is also supported by R-SamBada [219], an R software package that provides a complete pipeline for landscape genomic analyses, from the retrieval of environmental variables at sampling locations to gene annotation using the Ensembl genome browser.
Other landscape genomics software include BAYENV [220], which uses the Bayesian method to compute correlations between allele frequencies and ecological variables, taking into account differences in sample size and population structure; LFMM [211,221], which identifies gene-environment associations and SNPs with allele frequencies that correlate with clines of environmental variables; and SGLMM [222], which extends the BAYENV approach [223] by using a spatially explicit model and calculating inferences with an Integrated Nested Laplace Approximation and Stochastic Partial Differential Equation (SPDE). BayPass [224] builds on BAYENV to capture linkage disequilibrium information. BAYESCENV [225] produces an FST-based genome scan, taking into account environmental differences between populations. The latest version of LFMM [226] improves on both scalability and speed with respect to other GEA methods using a least-squares approach to estimate cofounders. Moreover, LFMM uses several categories of genomic data which are not restricted to genotypes.
Landscape genomics studies often use population genomics software (e.g., LOSITAN based on the FDist model [227,228]) to compare the sets of candidate loci obtained from different approaches: see BayeScan [229] and Bayenv [223]. A comparison of results allows for consolidation, as the accuracy of methods is known to differ (see, e.g., [213]). Samβada / R–SamBada [219] gives reliable results when the population structure is weak, while LFMM2 [226] is better suited to detect selection signatures in well-structured populations. Analyses of simulated data using, e.g., CDPOP [230] is usually advised to demonstrate the effectiveness of the method before moving to the analysis of empirical data (see, e.g., [211,213,231]). GEONOMICS, a Python package, performs forward-time, individual-based, continuous-space population genomic simulations on complex landscapes [232]. GEONOMICS includes several analytical steps using models of a landscape with one or more environmental layers (geotiff files as input), each of which can undergo environmental changes, as well as species having genomes with realistic architecture and associated phenotypes. Species undergo non-Wright–Fisher evolution in continuous space, with localized mating and mortality. The results produced are useful for a wide variety of theoretical and empirical purposes such as species conservation and management.

4.5. Artificial Intelligence and Machine Learning Approaches

With advances in genomic technology and more sophisticated sensing systems, “big data” sets are being created and a large amount of data needs to be stored every day [233]. These data sets will potentially reveal changes in genomes that adapt animals to a wide range of conditions and environments. However, the information is a mixture of homogeneous and heterogeneous data types where the relationships among parameters may be hidden or difficult to identify. Artificial Intelligence (AI) and Machine Learning (ML) methods are increasingly used to extract information from this type of data to overcome the limits of traditional linear models (250, 251) (see Box 6). ML and AI have not yet been fully applied to study adaptation to climate change in livestock; however, the role of big data and machine learning will become increasingly important for modern farming [234].
ML methods have been used in the quest for regions associated with adaptation, in particularly to detect de novo mutations and selective sweeps for previously segregating variants in humans [235]. The S/HIC Deep Learning (DL) model has shown that most human mutations are neutral in populations, and that those conferring an adaptive advantage only rise in frequency when a change in the environment gives advantages to individuals carrying a particular mutation [236]. This approach has been used to identify genes associated with metabolism in a southern African ethnic groups using the SWIF(r) DL algorithm [237]. Variants of these genes arose thousands of years ago to store fat when food was scarce.
There are a few examples of the use of ML in livestock genetics and breeding [196,238,239], and new DL genetic models are only just being tested [240,241,242,243]. The identification of SNPs directly associated with candidate genes affecting growth traits in Brahman cattle was more successful using ML Gradient Boosting Machine (GBM) than Random Forest statistical methods [241]. ML algorithms have been used together with RNA-Seq expression data to identify genes associated with feed efficiency in pigs, and to classify animals’ phenotypic extreme for residual feed intake [244].
Box 6. Artificial Intelligence and Machine Learning.
Artificial Intelligence (AI) uses algorithms that automate the decision process [245], while Machine Learning (ML) uses AI to automatically learn complex relationships and patterns in data [246,247]. ML algorithms may be unsupervised or supervised. The former explores the dataset structure without prior knowledge of data organization, while the latter uses prior knowledge to train the model and predict the outcome in a test dataset [248]. ML algorithms are adapted to explore nonlinear relationships [249]. Deep learning (DL) creates multiple processing layers (neural networks), which mimic the structure of a human brain, to extract information and learn from the input data. DL is being used to discover intricate structures in large datasets [246,250]. However, the neural network models are a “black box” as they are hidden as they develop. Tools are being developed to dissect the layers of the models developed to understand the neural network process; one example are the saliency maps [251,252].
ML methods mainly focus on prediction, while classical statistical methods rely on inference [253]. ML has been used to recognize the location of specific sequence elements (i.e., splice sites, promoters, etc.) and to combine genomic elements to identify and annotate genomic features, e.g., to identify UTR, introns, and exons, and to functionally annotate genes [235]. For example, S/HIC (https://github.com/kern-lab/shIC) is an ML classifier developed to detect targets of adaptive natural selection from whole genome sequencing data.
Efficient DL software tools such as Tensorflow and Keras Python libraries, and the availability of supercomputing using graphics processing unit technology (GPU), have opened the way to the integration of multi-omics big data with environmental variables.
Table 1. Software for genome-wide analyses.
Table 1. Software for genome-wide analyses.
SoftwareMethod ApplicationRef.Link
ArlequinTajima’s DSelection signatures[254]http://cmpg.unibe.ch/software/arlequin35/
BayeScanFSTSelection Signatures, Landscape genomics[229]http://cmpg.unibe.ch/software/BayeScan/
BcftoolsROHSelection signatures[255]https://github.com/samtools/bcftools
DnaSPTajima’s D and Fay and Wu’s statistic Selection signatures http://www.ub.edu/dnasp/
HapbinEHHSelection signatures[256]https://github.com/evotools/hapbin
hapFLKhapFLKSelection signatures[179]https://forge-dga.jouy.inra.fr/projects/hapflk
HierFstat (R package)FSTSelection signatures[257]https://cran.r-project.org/web/packages/hierfstat/index.html
KINGROHSelection signatures[258]https://www.kingrelatedness.com/
PLINKFST, ROHGWAS, Selection Signatures[259]https://www.cog-genomics.org/plink/2.0/https://www.cog-genomics.org/plink/
PopGenomeTajima’s DSelection signatures[260]https://cran.r-project.org/web/packages/PopGenome/index.html
PoPoolationTajima’s DSelection signatures[261]https://sourceforge.net/p/popoolation/wiki/Main/
rehh (R package)EHHSelection signatures[262]https://cran.r-project.org/web/packages/rehh/index.html
SelscanEHHSelection signatures[263]https://github.com/szpiech/selscan
VariScanTajima’s DSelection signatures[264]http://www.ub.edu/softevol/variscan/
VCFtoolsFST, Tajima’s DSelection signatures[265]http://vcftools.sourceforge.net/
EMMAXGWAS based on variance component modelGWAS[266]http://genetics.cs.ucla.edu/emmax
GCTAGWAS based on genome-wide complex trait analysisGWAS[267]http://gump.qimr.edu.au/gcta
BayesRBayesian mixture modelGWAS[268]http://www.cnsgenomics.com/software/
MatSAMLogistic regressionLandscape genomics[208]www.econogene.eu/software/sam/
Samβada, R.SamBada (R package)GEA based on logistic regression/spatial autocorrelationLandscape genomics[213,219]https://github.com/Sylvie/sambada/releases/tag/v0.8.3https://cran.r-project.org/package=R.SamBada
BAYENVGEA based on Bayesian regressionLandscape genomics[220]https://gcbias.org/bayenv/
LFMM2 (R package)GEA based on latent factor mixed modelsLandscape genomics[221,226]https://bcm-uga.github.io/lfmm/
SGLMMGEA based on allele-environment association analysisLandscape genomics[222]-
BayPassGEA corrected for the covariance structure among the population allele frequenciesLandscape genomics[224]http://www1.montpellier.inra.fr/CBGP/software/baypass/
BAYESCENVGEA based on FST genome-scanLandscape genomics[225]https://github.com/devillemereuil/bayescenv
LOSITANFST Landscape genomics[227]https://mybiosoftware.com/lositan-1-0-0-selection-detection-workbench.html
PCAdmixSupervised LAI Local Ancestry Inference[186]https://sites.google.com/site/pcadmix/home
TractorLA-aware regression model Local Ancestry Inference[187]https://github.com/eatkinson/Tractor
LAMPLAI accounting for recombinationLocal Ancestry Inference[188]http://lamp.icsi.berkeley.edu/lamp/
MOSAIC (R package)Unsupervised LAILocal Ancestry Inference[193]https://maths.ucd.ie/~mst/MOSAIC/
RFMixLAI based on conditional random fieldLocal Ancestry Inference[194]https://github.com/slowkoni/rfmix
LoterLAI for species other than humansLocal Ancestry Inference[195]https://github.com/bcm-uga/Loter
GHap (R package)Unsupervised LAILocal Ancestry Inference[196] https://cran.r-project.org/package=GHap
PSIKO2Unsupervised LAILocal Ancestry Inference[197]https://www.uea.ac.uk/computing/psiko
SWIF(r)Probabilistic method to detect selective sweepsDeep Learning[237]https://github.com/ramachandran-lab/SWIFr

5. Conclusions

To maintain animal welfare and as a consequence productivity and production efficiency, breeds have to be well adapted to the environmental conditions in which they are kept. Rapid climate change inevitably calls for the use of various countermeasures to manage animals appropriately. Temperature mitigation methods (shaded area, water wetting, ventilation, air conditioning) are possible solutions; however, these can only be used when animals are kept in shelters and are not applicable to range-type farming systems. Most structural solutions to control the environment of animals have a high cost, and many have energy requirements that further contribute to climate change. Therefore, addressing livestock adaptation by breeding animals that are intrinsically more tolerant to extreme conditions is a more sustainable solution. Decreasing stress and increasing animal welfare is important for farmers and the general public. Animals stressed by high temperatures may be less able to cope with other stressors such as pollutants, dust, restraint, social mixing, transport, etc., that further affect welfare and productivity. Innovation in sensors and linking these into the “internet of things” (IoT) to collect and exchange data is increasing our ability to record environmental variables and animal welfare status and provide input to systems dedicated to the control of environmental conditions and provision of early warning of discomfort in individual animals. In the longer term, collecting such data will contribute to understanding the genetics underpinning tolerance and adaptation to environmental and other stressors in order to select animals better suited to different conditions. The resulting increase in efficiency will have additional benefits in terms of reducing greenhouse gas emissions, particularly methane from ruminants, which currently make a significant contribution to climate change.
Breed substitution by introducing breeds known to have particular resilience, e.g., to drought, temperature extremes or disease, may be a solution. This approach would facilitate a rapid response to climate change, although it is not ideal as breeds more tolerant of hot climates generally have low productivity. Additionally, imported breeds may not adapt to local conditions such as available feed resources and disease challenge.
Crossbreeding between highly productive and heat tolerant breeds is an approach that is currently used in tropical areas including Australia, the southern USA and Brazil, where crossing productive taurine breeds with heat adapted indicine breeds facilitates improved production in extreme conditions. Selection of these cross-bred populations has produced stable breeds that show good productivity and adaptation, such as the Brangus from the USA [269] and the Australian Droughtmaster [270]. O’Neil et al. [271] have reviewed the use of crossbred lines in tropical high tick challenge areas of Australia. However, crossbreeding programs should be properly planned, organised and monitored, as indiscriminate crosses may cause the genetic erosion of local breeds and the loss of their adaptation.
Accelerated selection for thermal tolerance and resilience to new endemic diseases is also a possible sustainable solution. In this case, genomics plays a key role together with phenotype recording and the collection of epidemiological and environmental data. Research is approaching the challenging task of identifying genes having adaptive value using a range of methods, including those described in this review. Specific variants of major genes exist in local genetic resources, as demonstrated by the SLICK mutation associated with heat tolerance. However, identifying causal genes and variants is difficult, requiring large data sets which are often not available or affordable for livestock, and a focused effort to refine and test candidate genes. Therefore, most studies have simply localized genetic effects to chromosomal regions or quantitative trait loci (QTL) in genome-wide association studies. Additionally, it is now clear that most adaptation traits have complex genetic control, making the genetic basis difficult to unravel. Nevertheless, markers having significant effects can be used in selection programmes using marker assisted selection or by weighting particular SNPs within QTL regions in genomic selection estimates. Although genomics is presently only scratching the surface of the control mechanism of these traits, comparison between methods, studies, breeds and even species is starting to reveal that morphology, energy and lipid metabolism, and the immune system are key factors in adaptation, with some genes being consistently identified as carrying variants modulating adaptation. The identification of these genes confirms the importance of the conservation of local genetic resources as reservoirs of useful alleles. The evaluation and improvement of these breeds or the transfer of adaptive variants into highly selected breeds are the next steps to better match livestock to harsh conditions while maintaining productivity. These steps may be accelerated by marker-assisted or genomic selection, and even more rapidly by novel tools such as gene editing where such approaches are socially accepted. Parallel breeding for adaptation to climate change and the mitigation of the impact of livestock on climate change is probably the hardest challenge that the livestock sector has ever faced, but it is now urgent. The challenge can only be won if research, industry, decision makers and funders join forces with the objective of satisfying the rights future generations to a healthy diet and a clean planet.

Author Contributions

Conceptualization, P.A.-M. and J.L.W.; writing—original draft preparation, P.A.-M., J.L.W., G.C., L.C., M.M., R.N., M.B., S.J., E.V., M.S., E.S., M.M.P.; writing—review and editing, P.A.-M. and J.L.W.; supervision, P.A.-M. and J.L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the following projects: HIGHLANDER (HIGH performance computing to support smart LAND sERvices) project, which is funded by the Connecting European Facility (CEF) Telecommunications sector under agreement number INEA/CEF/ICT/A2018/1815462; SEBASTIEN (Smarter livEstock Breeding through Advanced Services Tailoring Innovative and multi-sourcE data to users’ Needs) project, which is funded by the Connecting European Facility (CEF) Telecommunications sector under agreement number INEA/CEF/ICT/A2020/2373580; IMAGE, Innovative Management of Animal Genetic Resources (IMAGE) project, which received funding from the European Union’s Horizon 2020 Research and Innovation Programme Grant 677353; SCALA-MEDI, Improving sustainability and quality of Sheep and Chicken production by leveraging the Adaptation potential of LocAl breeds in the MEDIterranean area, funded by the European Union’s Horizon 2020 Partnership for Research and Innovation in the Mediterranean Area Programme (PRIMA) Grant 2012; Departments of Excellence-2018 Program (Dipartimenti di Eccellenza) of the Italian Ministry of Education, University and Research, DIBAF-Department of University of Tuscia, Project “Landscape 4.0–food, wellbeing and environment”. LEO, Livestock Environment Opendata, 16.2–PSRN 2014–2020 financed through Fondo Europeo Agricolo per lo Sviluppo Rurale (FEASR). NOVALSELPROV, InNOVAzioni nelle PROduzioni casearie OVine di sicilia e sardegna, PON project n. ARS01_00580. The authors wish to thank Fondazione Romeo ed Enrica Invernizzi. This study was supported by the Doctoral School on the Agro-Food System (Agrisystem) of the Università Cattolica del Sacro Cuore (Italy).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study, in the collection, analyses, or interpretation of data, in the writing of the manuscript, or in the decision to publish the results.

References

  1. Masson-Delmotte, V.; Zhai, P.; Pörtner, H.O.; Roberts, D.; Skea, J.; Shukla, P.R.; Pirani, A.; Moufouma-Okia, W.; Péan, C.; Pidcock, R.; et al. IPCC, 2018: Global Warming of 1.5 °C. An IPCC Special Report on the Impacts of Global Warming of 1.5 °C above Pre-Industrial Levels and Related Global Greenhouse Gas Emission Pathways, in the Context of Strengthening the Global Response to the Threat of Climate change, sustainable development, and efforts to eradicate poverty. 2018, in press. [Google Scholar]
  2. Ames, D.R.; Insley, L.W. Wind-Chill Effect for Cattle and Sheep. J. Anim. Sci. 1975, 40, 161–165. [Google Scholar] [CrossRef] [Green Version]
  3. Toghiani, S.; Hay, E.H.; Roberts, A.; Rekaya, R. Impact of Cold Stress on Birth and Weaning Weight in a Composite Beef Cattle Breed. Livest. Sci. 2020, 236, 104053. [Google Scholar] [CrossRef]
  4. Pezzopane, R.M.; Carlos, A.; Bernardi, C.; Azenha, M.V.; Anch, P.P.; Bosi, C.; Pedroso, F.; Esteves, N. Production and Nutritive Value of Pastures in Integrated Livestock Production Systems: Shading and Management Effects. Sci. Agric. 2020, 77, 1–9. [Google Scholar] [CrossRef] [Green Version]
  5. Belasco, E.; Cheng, Y.; Schroeder, T.C. The Impact of Extreme Weather on Cattle Feeding Profits. J. Agric. Resour. Econ. 2015, 40, 285–305. [Google Scholar]
  6. Lees, A.M.; Sejian, V.; Wallage, A.L.; Steel, C.C.; Mader, T.L.; Lees, J.C.; Gaughan, J.B. The Impact of Heat Load on Cattle. Animals 2019, 9, 322. [Google Scholar] [CrossRef] [Green Version]
  7. Rust, W.; Holman, I.; Bloomfield, J.; Cuthbert, M.; Corstanje, R. Understanding the Potential of Climate Teleconnections to Project Future Groundwater Drought. Hydrol. Earth Syst. Sci. 2019, 23, 3233–3245. [Google Scholar] [CrossRef] [Green Version]
  8. Thornton, P.K.; van de Steeg, J.; Notenbaert, A.; Herrero, M. The Impacts of Climate Change on Livestock and Livestock Systems in Developing Countries: A Review of What We Know and What We Need to Know. Agric. Syst. 2009, 101, 113–127. [Google Scholar] [CrossRef]
  9. Rotter, R.; Van de Geijn, S.C. Climate Change Effects On Plant Growth, Crop Yield and Livestock. Clim. Chang. 1999, 43, 651–681. [Google Scholar] [CrossRef]
  10. Wheeler, T.; Reynolds, C. Predicting the Risks from Climate Change to Forage and Crop Production for Animal Feed. Anim. Front. 2013, 3, 36–41. [Google Scholar] [CrossRef] [Green Version]
  11. Ortiz-Bobea, A.; Knippenberg, E.; Chambers, R.G. Growing Climatic Sensitivity of U.S. Agriculture Linked to Technological Change and Regional Specialization. Sci. Adv. 2018, 4, eaat4343. [Google Scholar] [CrossRef] [Green Version]
  12. Gaughan, J.B.; Cawdell-Smith, A.J. Impact of climate change on livestock production and reproduction. In Climate Change Impacts on Livestock: Adaptation and Mitigation; Sejian, V., Gaughan, J., Baumgard, L., Prasad, C., Eds.; Springer India: New Delhi, India, 2017; pp. 51–60. [Google Scholar]
  13. Hahn, G.; Mader, T.; Spiers, D.; Gaughan, J.; Nienaber, J.; Eigenberg, R.; Brown-Brandl, T.; Hu, Q.; Griffin, D.; Hugenford, L. Heat Wave Impacts on Feedlot Cattle: Considerations for Improved Environmental Management. In Proceedings of the Livestock Environment International Symposium Proceedings, Louisville, KY, USA, 21–23 May 2001. [Google Scholar]
  14. Mader, T.; Davis, S.; Gaughan, J.; Brown-brandl, T. Wind Speed and Solar Radiation Adjustments for the Temperature-Humidity Index. In Proceedings of the 16th Conference on Biometeorology and Aerobiology, Vancouver, BC, Canada, 23–27 August 2004; pp. 1–6. [Google Scholar]
  15. National Research Council Effect of enviroment on utrient requirement of domestic animals. In Subcommittee on Environmental Stress; National Research Council (Ed.) The National Academies Press: Washington, DC, USA, 1981. [Google Scholar]
  16. Hahn, G. Global Warming and Potential Impacts on Cattle and Swine in Tropical and Temperate Areas. In Proceedings of the Brazilian Congress of Biometeorology, USDA, Ljubljana, Slovenia, 1–8 September 1996. [Google Scholar]
  17. Thom, E.C. The Discomfort Index. Weatherwise 1959, 12, 57–61. [Google Scholar] [CrossRef]
  18. Amundson, J.L.; Mader, T.L.; Rasby, R.J.; Hu, Q.S. Environmental Effects on Pregnancy Rate in Beef Cattle 1. J. Anim. Sci. 2006, 84, 3415–3420. [Google Scholar] [CrossRef] [Green Version]
  19. Shukla, P.R.; Skea, J.; Buendia, E.C.; Masson-Delmotte, V.; Pörtner, H.-O.; Roberts, D.C.; Zhai, P.; Slade, R.; Connors, S.; van Diemen, R.; et al. IPCC, 2019: Climate Change and Land: An IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial Ecosystems. 2019, in press. [Google Scholar]
  20. Lean, J.L.; Rind, D.H. How Will Earth’s Surface Temperature Change in Future Decades? Geophys. Res. Lett. 2009, 36, 1–5. [Google Scholar] [CrossRef] [Green Version]
  21. Chapman, S.C.; Chakraborty, S.; Dreccer, F.; Howden, M. Plant Adaptation to Climate Change—Opportunities and Priorities in Breeding. Crop Pasture Sci. 2012, 63, 251–268. [Google Scholar] [CrossRef] [Green Version]
  22. Calleja-Cabrera, J.; Boter, M.; Oñate-Sánchez, L.; Pernas, M. Root Growth Adaptation to Climate Change in Crops. Front. Plant Sci. 2020, 11, 544. [Google Scholar] [CrossRef]
  23. Tubiello, F.N.; Soussana, J.F.; Howden, S.M. Crop and Pasture Response to Climate Change. Proc. Natl. Acad. Sci. USA 2007, 104, 19686–19690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Hidosa, D.; Guyo, M. Journal of Fisheries & Climate Change Effects on Livestock Feed Resources: A Review. J. Fish. Livest. Prod. 2017, 5, 5–8. [Google Scholar] [CrossRef] [Green Version]
  25. Morton, J.F. The Impact of Climate Change on Smallholder and Subsistence Agriculture. Proc. Natl. Acad. Sci. USA 2007, 104, 19680–19685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. O’Mara, F.P. The Role of Grasslands in Food Security and Climate Change. Ann. Bot. 2012, 110, 1263–1270. [Google Scholar] [CrossRef] [Green Version]
  27. Baylis, M.; Risley, C. Infectious Diseases, Climate Change Effects on. In Infectious Diseases; Kanki, P., Grimes, D., Eds.; Springer: New York, NY, USA, 2013; pp. 117–146. [Google Scholar]
  28. Baylis, M.; Githeko, A.K. The Effects of Climate Change on Infectious Diseases of Animals; Report for the Foresight Project on Detection of Infectious Diseases; Department of Trade and Industry, UK Government: London, UK, 2006.
  29. Wilson, A.J.; Mellor, P.S. Bluetongue in Europe: Past, Present and Future. Philos. Trans. R. Soc. London. B 2009, 364, 2669–2681. [Google Scholar] [CrossRef]
  30. Olwoch, J.M.; Reyers, B.; Engelbrecht, F.A.; Erasmus, B.F.N. Climate Change and the Tick-Borne Disease, Theileriosis (East Coast Fever) in Sub-Saharan Africa. J. Arid Environ. 2008, 72, 108–120. [Google Scholar] [CrossRef]
  31. Kenyon, F.; Sargison, N.D.; Skuce, P.J.; Jackson, F. Sheep Helminth Parasitic Disease in South Eastern Scotland Arising as a Possible Consequence of Climate Change. Vet. Parasitol. 2009, 163, 293–297. [Google Scholar] [CrossRef] [PubMed]
  32. Fox, N.J.; Glenn, M.; Davidson, R.S.; White, P.C.L.; Hutchings, M. Climate-Driven Tipping-Points Could Lead to Sudden, High-Intensity Parasite Outbreaks. R. Soc. Open Sci. 2015, 2, 2–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Lau, C.L.; Smythe, L.D.; Craig, S.B.; Weinstein, P. Climate Change, Flooding, Urbanisation and Leptospirosis: Fuelling the Fire? Trans. R. Soc. Trop. Med. Hyg. 2010, 104, 631–638. [Google Scholar] [CrossRef] [PubMed]
  34. Rojas-Downing, M.M.; Nejadhashemi, A.P.; Harrigan, T.; Woznicki, S.A. Climate Change and Livestock: Impacts, Adaptation, and Mitigation. Clim. Risk Manag. 2017, 16, 145–163. [Google Scholar] [CrossRef]
  35. Buontempo, C.; Hutjes, R.; Beavis, P.; Berckmans, J.; Cagnazzo, C.; Vamborg, F.; Bergeron, C.; Almond, S.; Amici, A.; Ramasamy, S.; et al. Fostering the Development of Climate Services through Copernicus Climate Change Service (C3S) for Agriculture Applications. Weather Clim. Extrem. 2020, 27, 1–8. [Google Scholar] [CrossRef]
  36. Fick, S.E.; Hijmans, R.J. WorldClim 2: New 1-Km Spatial Resolution Climate Surfaces for Global Land Areas. Int. J. Climatol. 2017, 37, 4302–4315. [Google Scholar] [CrossRef]
  37. Karger, D.N.; Conrad, O.; Böhner, J.; Kawohl, T.; Kreft, H.; Soria-auza, R.W.; Zimmermann, N.E.; Linder, H.P.; Kessler, M. Data Descriptor: Climatologies at High Resolution for the Earth’s Land Surface Areas. Sci. Data 2017, 4, 1–20. [Google Scholar] [CrossRef] [Green Version]
  38. Kriticos, D.J.; Webber, B.L.; Leriche, A.; Ota, N.; Macadam, I.; Bathols, J.; Scott, J.K. CliMond: Global High-Resolution Historical and Future Scenario Climate Surfaces for Bioclimatic Modelling. Methods Ecol. Evol. 2012, 3, 53–64. [Google Scholar] [CrossRef]
  39. Lima-Ribeiro, M.S.; Varela, S.; González-Hernández, J.; de Oliveira, G.; Diniz-Filho JA, F.; Terribile, L.C. EcoClimate: A Database of Climate Data from Multiple Models for Past, Present, and Future for Macroecologists and Biogeographers. Biodivers. Inform. 2015, 10, 1–21. [Google Scholar] [CrossRef] [Green Version]
  40. Title, P.O.; Bemmels, B. ENVIREM: An Expanded Set of Bioclimatic and Topographic Variables Increases Flexibility and Improves Performance of Ecological Niche Modeling. Ecography 2018, 41, 291–307. [Google Scholar] [CrossRef] [Green Version]
  41. Vega, G.C.; Pertierra, L.R.; Olalla-tárraga, M.Á. Data Descriptor: MERRAclim, a High-Resolution Global Dataset of Remotely Sensed Bioclimatic Variables for Ecological Modelling. Sci. Data 2017, 4, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Noce, S.; Caporaso, L.; Santini, M. A New Global Dataset of Bioclimatic Indicators. Sci. Data 2020, 7, 1–12. [Google Scholar] [CrossRef]
  43. Cui, D.; Liang, S.; Wang, D.; Liu, Z. Köppen-Geiger Climate Classification and Bioclimatic Variables. Earth Syst. Sci. Data 2021, 1–34. [Google Scholar] [CrossRef]
  44. Thépaut, J.; Dee, D.; Engelen, R.; Pinty, B. The Copernicus Programme and Its Climate Change Service. In Proceedings of the IGARSS 2018—2018 IEEE International Geoscience and Remote Sensing Symposium, Valencia, Spain, 22–27 July 2018; pp. 1591–1593. [Google Scholar]
  45. Sillmann, J.; Kharin, V.V.; Zhang, X.; Zwiers, F.W.; Bronaugh, D. Climate Extremes Indices in the CMIP5 Multimodel Ensemble: Part 1. Model Evaluation in the Present Climate. J. Geophys. Res. 2013, 118, 1716–1733. [Google Scholar] [CrossRef]
  46. Sillmann, J.; Kharin, V.V.; Zwiers, F.W.; Zhang, X.; Bronaugh, D. Climate Extremes Indices in the CMIP5 Multimodel Ensemble: Part 2. Future Climate Projections. J. Geophys. Res. 2013, 118, 2473–2493. [Google Scholar] [CrossRef]
  47. Naderi, S.; Rezaei, H.R.; Taberlet, P.; Zundel, S.; Rafat, S.; Naghash, H.; El-Barody, M.A.A.; Ertugrul, O.; Pompanon, F. Large-Scale Mitochondrial DNA Analysis of the Domestic Goat Reveals Six Haplogroups with High Diversity. PLoS ONE 2007, 2, e1012. [Google Scholar] [CrossRef] [Green Version]
  48. Zeder, M.A. Domestication and Early Agriculture in the Mediterranean Basin: Origins, Diffusion, and Impact. Proc. Natl. Acad. Sci. USA 2008, 19, 11597–11604. [Google Scholar] [CrossRef] [Green Version]
  49. Zeder, M.A. Core Questions in Domestication Research. Proc. Natl. Acad. Sci. USA 2015, 17, 3192–3198. [Google Scholar] [CrossRef] [Green Version]
  50. Legge, T. The beginnings of caprine domestication. In The Origins and Spread of Agriculture and Pastoralism in Eurasia; Harris, D.R., Ed.; Smithsonian Institution Press: New York, NY, USA, 1996; pp. 238–262. [Google Scholar]
  51. Vigne, J. The Origins of Animal Domestication and Husbandry: A Major Change in the History of Humanity and the Biosphere. Comptes Rendus Biol. 2011, 334, 171–181. [Google Scholar] [CrossRef]
  52. Mchugo, G.P.; Dover, M.J.; Machugh, D.E. Unlocking the Origins and Biology of Domestic Animals Using Ancient DNA and Paleogenomics. BMC Biol. 2019, 17, 98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Park, S.D.E.; Magee, D.A.; McGettigan, P.A.; Teasdale, M.D.; Edwards, C.J.; Lohan, A.J.; Murphy, A.; Braud, M.; Donoghue, M.T.; Liu, Y.; et al. Genome Sequencing of the Extinct Eurasian Wild Aurochs, Bos Primigenius, Illuminates the Phylogeography and Evolution of Cattle. Genome Biol. 2015, 16, 234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Daly, K.G.; Delser, P.M.; Mullin, V.E.; Scheu, A.; Mattiangeli, V.; Teasdale, M.D.; Hare, A.J.; Burger, J.; Verdugo, M.P.; Collins, M.J.; et al. Ancient Goat Genomes Reveal Mosaic Domestication in the Fertile Crescent. Science 2018, 361, 24–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Verdugo, M.P.; Mullin, V.E.; Scheu, A.; Mattiangeli, V.; Daly, K.G.; Delser, P.M.; Hare, A.J.; Burger, J.; Collins, M.J.; Kehati, R.; et al. Ancient Cattle Genomics, Origins and Rapid Turnover in the Fertile Crescent. Science 2019, 176, 173–176. [Google Scholar]
  56. Zheng, Z.; Wang, X.; Li, M.; Li, Y.; Yang, Z.; Wang, X.; Pan, X.; Gong, M.; Zhang, Y.; Guo, Y.; et al. The Origin of Domestication Genes in Goats. Sci. Adv. 2020, 6, eaaz5216. [Google Scholar] [CrossRef] [PubMed]
  57. Librado, P.; Gamba, C.; Gaunitz, C.; Der Sarkissian, C.; Pruvost, M.; Albrechtsen, A.; Fages, A.; Khan, N.; Schubert, M.; Jagannathan, V.; et al. Ancient Genomic Changes Associated with Domestication of the Horse. Science 2017, 356, 442–445. [Google Scholar] [CrossRef] [PubMed]
  58. Pendleton, A.; Shen, F.; Taravella, A.; Emery, S.; Veeramah, K.; Boyko, A.; Kidd, J.M. Comparison of Village Dog and Wolf Ge-Nomes Highlights the Role of the Neural Crest in Dog Domestication. BMC Biol. 2018, 16, 64. [Google Scholar] [CrossRef] [Green Version]
  59. Hannemann, F.; Bichet, A.; Ewen, K.M.; Bernhardt, R. Cytochrome P450 Systems—Biological Variations of Electron Transport Chains. Biochim. Biophys. Acta—Gen. Subj. 2007, 1770, 330–344. [Google Scholar] [CrossRef]
  60. Weinberg, P.J. Capra Cylindricornis. Mamm. Species 2002, 695, 1–9. [Google Scholar] [CrossRef]
  61. Liu, Y.; Weyrich, L.; Llamas, B. More Arrows in the Ancient DNA Quiver: Use of Paleoepigenomes and Paleomicrobiomes to Investigate Animal Adaptation to Environment. Mol. Biol. Evol. 2020, 37, 307–319. [Google Scholar] [CrossRef] [PubMed]
  62. Collier, R.J.; Baumgard, L.H.; Zimbelman, R.B.; Xiao, Y. Heat Stress: Physiology of Acclimation and Adaptation. Anim. Front. 2019, 9, 12–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Gaughan, J.B.; Sejian, V.; Mader, T.L.; Dunshea, F.R. Adaptation Strategies: Ruminants. Anim. Front. 2019, 9, 47–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Mignon-grasteau, S.; Boissy, A.; Bouix, J.; Faure, J.; Fisher, A.D.; Hinch, G.N.; Jensen, P.; Le, P.; Prunet, P.; Vandeputte, M.; et al. Genetics of Adaptation and Domestication in Livestock B. Livest. Prod. Sci. 2005, 93, 3–14. [Google Scholar] [CrossRef]
  65. Galal, S.; Rasoul, F.A.; Annous, M.R.; Shoat, I. Small Ruminant Breeds of Egypt; Aleppo, 2005; International Center for Agricultural Research in Dry Areas (ICARDA): Aleppo, Syria, 2005; 196p. [Google Scholar]
  66. Mazzullo, G.; Rifici, C.; Cammarata, F.; Caccamo, G.; Rizzo, M.; Piccione, G.; Practician, V.; Practicioner, V. Effect of Different Environmental Conditions on Some Haematological Parameters in Cow. Ann. Anim. Sci. 2014, 14, 947–954. [Google Scholar] [CrossRef] [Green Version]
  67. Hansen, P. Physiological and Cellular Adaptations of Zebu Cattle to Thermal Stress. Anim. Reprod. Sci. 2004, 82–83, 349–360. [Google Scholar] [CrossRef]
  68. Utsunomiya, Y.T.; Milanesi, M.; Fortes, M.R.S. Genomic Clues of the Evolutionary History of Bos Indicus Cattle. Anim. Genet. 2019, 50, 557–568. [Google Scholar] [CrossRef] [Green Version]
  69. Trail, J.C.M.; Gregory, K.E.; Marples, H.J.S.; Kakonge, J. Comparison of Bos Taurus-Bos Indicus Breed Crosses with Straightbred Bos Indicus Breeds of Cattle for Maternal and Individual Traits. J. Anim. Sci. 1985, 60, 1181–1187. [Google Scholar] [CrossRef]
  70. Madalena, F.E. DAIRY ANIMALS|Bos Indicus Breeds and Bos Indicus × Bos Taurus Crosses. Encycl. Dairy Sci. 2002, 576–585. [Google Scholar] [CrossRef]
  71. Hill, D.; Wall, E. Dairy Cattle in a Temperate Climate: The Effects of Weather on Milk Yield and Composition Depend on Management. Animal 2015, 9, 138–149. [Google Scholar] [CrossRef] [Green Version]
  72. da Silva, V.P.R.; e Silva, R.A.; Cavalcanti, E.P.; Braga, C.C.; de Azevedo, P.V.; Singh, V.P.; Pereira, E.R.R. Trends in Solar Radiation in NCEP/NCAR Database and Measurements in Northeastern Brazil. Sol. Energy 2010, 84, 1852–1862. [Google Scholar] [CrossRef]
  73. Herbut, P.; Angrecka, S.; Walczak, J. Environmental Parameters to Assessing of Heat Stress in Dairy Cattle—A Review. Int. J. Biometeorol. 2018, 62, 2089–2097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. St-Pierre, N.R.; Cobanov, B.; Schnitkey, G. Economic Losses from Heat Stress by US Livestock Industries. J. Dairy Sci. 2003, 86, E52–E77. [Google Scholar] [CrossRef] [Green Version]
  75. Baêta, F.C.; Meador, N.; Shanklin, M.D.; Johnson, H.D. Equivalent Temperature Index at Temperatures above the Thermoneu-Tral for Lactating Dairy Cows. In Proceedings of the Meeting of the American Society of Agricultural engineers, Chicago, IL, USA, 1987; Available online: https://agris.fao.org/agris-search/search.do?recordID=US8853966 (accessed on 21 July 2021).
  76. Mader, T.L.; Davis, M.S.; Brown-Brandl, T. Environmental Factors Influencing Heat Stress in Feedlot Cattle. J. Anim. Sci. 2006, 84, 712–719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Eigenberg, R.A.; Brown-Brandl, T.M.; Nienaber, J.A.; Hahn, G.L. Dynamic Response Indicators of Heat Stress in Shaded and Non-Shaded Feedlot Cattle, Part 2: Predictive Relationships. Biosyst. Eng. 2005, 91, 111–118. [Google Scholar] [CrossRef] [Green Version]
  78. Rashamol, V.P.; Sejian, V.; Pragna, P.; Lees, A.M.; Bagath, M.; Krishnan, G.; Gaughan, J.B. Prediction Models, Assessment Methodologies and Biotechnological Tools to Quantify Heat Stress Response in Ruminant Livestock. Int. J. Biometeorol. 2019, 63, 1265–1281. [Google Scholar] [CrossRef] [Green Version]
  79. Gaughan, J.B.; Mader, T.L.; Holt, S.M.; Lisle, A. A New Heat Load Index for Feedlot Cattle. J. Anim. Sci. 2008, 86, 226–233. [Google Scholar] [CrossRef] [Green Version]
  80. Mader, T.L.; Johnson, L.J.; Gaughan, J.B.; Mader, T.L.; Johnson, L.J.; Gaughan, J.B. A Comprehensive Index for Assessing Environmental Stress in Animals. J. Anim. Sci. 2010, 88, 2153–2165. [Google Scholar] [CrossRef] [Green Version]
  81. Flori, L.; Gonzatti, M.I.; Thevenon, S.; Chantal, I.; Pinto, J.; Berthier, D.; Aso, P.M.; Gautier, M. A Quasi-Exclusive European Ancestry in the Senepol Tropical Cattle Breed Highlights the Importance of the Slick Locus in Tropical Adaptation. PLoS ONE 2012, 7, e36133. [Google Scholar] [CrossRef]
  82. Olson, T.A.; Lucena, C.; Chase, C.C.J.; Hammond, A.C. Evidence of a Major Gene Influencing Hair Length and Heat Tolerance in Bos Taurus Cattle. J. Anim. Sci. 2003, 81, 80–90. [Google Scholar] [CrossRef]
  83. Dikmen, S.; Alava, E.; Pontes, E.; Fear, J.M.; Dikmen, B.Y.; Olson, T.A.; Hansen, P.J. Differences in Thermoregulatory Ability Between Slick-Haired and Wild-Type Lactating Holstein Cows in Response to Acute Heat Stress. J. Dairy Sci. 2008, 91, 3395–3402. [Google Scholar] [CrossRef] [PubMed]
  84. Mariasegaram, M.; Chase, C.C.J.; Chaparro, J.X.; Olson, T.A.; Brenneman, R.A.; Niedz, R.P. The Slick Hair Coat Locus Maps to Chromosome 20 in Senepol-Derived Cattle. Anim. Genet. 2007, 38, 54–59. [Google Scholar] [CrossRef] [PubMed]
  85. Littlejohn, M.D.; Henty, K.M.; Tiplady, K.; Johnson, T.; Harland, C.; Lopdell, T.; Sherlock, R.G.; Li, W.; Lukefahr, S.D.; Shanks, B.C.; et al. Functionally Reciprocal Mutations of the Prolactin Signalling Pathway Define Hairy and Slick Cattle. Nat. Commun. 2014, 5, 5861. [Google Scholar] [CrossRef] [Green Version]
  86. Martinez, A.M.; Gama, L.T.; Canon, J.; Ginja, C.; Delgado, J.V.; Dunner, S.; Al, E. Genetic Footprints of Iberian Cattle in America 500 Years after the Arrival of Columbus. PLoS ONE 2012, 7, e49066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Porto-neto, L.R.; Bickhart, D.M.; Landaeta-hernandez, A.J.; Utsunomiya, Y.T.; Pagan, M.; Jimenez, E.; Hansen, P.J.; Dikmen, S.; Schroeder, S.G.; Kim, E.; et al. Convergent Evolution of Slick Coat in Cattle through Truncation Mutations in the Prolactin Receptor. Front. Genet. 2018, 9, 57. [Google Scholar] [CrossRef] [Green Version]
  88. Mundel, T.; Bunn, S.J.; Hooper, P.L.; Jones, D.A. The Effects of Face Cooling during Hyperthermic Exercise in Man: Evidence for an Integrated Thermal, Neuroendocrine and Behavioural Response. Exp. Physiol. 2007, 92, 187–195. [Google Scholar] [CrossRef]
  89. Dikmen, S.; Khan, F.A.; Huson, H.J.; Sonstegard, T.S.; Moss, J.I.; Dahl, G.E.; Hansen, P.J. The SLICK Hair Locus Derived from Senepol Cattle Confers Thermotolerance to Intensively Managed Lactating Holstein Cows. J. Dairy Sci. 2014, 97, 5508–5520. [Google Scholar] [CrossRef] [Green Version]
  90. Bahbahani, H.; Tijjani, A.; Mukasa, C.; Wragg, D. Signatures of Selection for Environmental Adaptation and Zebu × Taurine Hybrid Fitness in East African Shorthorn Zebu. Front. Genet. 2017, 8, 68. [Google Scholar] [CrossRef] [Green Version]
  91. Lander, E.S.; Linton, L.M.; Birren, B.; Nusbaum, C.; Zody, M.C.; Baldwin, J.; Devon, K.; Dewar, K.; Doyle, M.; FitzHugh, W.; et al. International Human Genome Consortium Initial Sequencing and Analysis of the Human Genome. Nature 2001, 409, 860–921. [Google Scholar]
  92. International Chicken Genome Sequencing Consortium. Sequence and Comparative Analysis of the Chicken Genome Provide Unique Perspectives on Vertebrate Evolution. Nature 2004, 432, 695–777. [Google Scholar] [CrossRef]
  93. The Bovine genome Sequencing and Analysis Consortium. The Genome Sequence of Taurine Cattle: A Window to Ruminant Biology and Evolution. Science 2009, 324, 522–529. [Google Scholar] [CrossRef] [Green Version]
  94. Groenen, M.A.M.; Archibald, A.; Uenishi, H.; Tuggle, C.K.; Takeuchi, Y.; Rothschild, M.F.; Rogel-Gailard, C.; Park, C.; Milan, D.; Megens, H.; et al. Analyses of Pig Genomes Provide Insight into Porcine Demography and Evolution. Nature 2012, 491, 393–398. [Google Scholar] [CrossRef]
  95. Jiang, Y.; Xie, M.; Chen, W.; Talbot, R.; Maddox, J.; Faraut, T.; Wu, C.; Muzny, D.M.; Li, Y.; Zhang, W.; et al. The Sheep Genome Illuminates Biology of the Rumen and Lipid Metabolism. Science 2014, 344, 1168–1173. [Google Scholar] [CrossRef] [Green Version]
  96. Dong, Y.; Xie, M.; Jiang, Y.; Xiao, N.; Du, X.; Zhang, W.; Tosser-klopp, G.; Wang, J.; Yang, S.; Liang, J.; et al. Sequencing and Automated Whole-Genome Optical Mapping of the Genome of a Domestic Goat (Capra hircus). Nat. Biotechnol. 2013, 31, 135–143. [Google Scholar] [CrossRef] [Green Version]
  97. Metzker, M.L. Sequencing Technologies—The next Generation. Nat. Rev. Genet. 2010, 11, 31–46. [Google Scholar] [CrossRef] [Green Version]
  98. Berry, D.P.; Kearney, J.F. Imputation of Genotypes from Low- to High-Density Genotyping Platforms and Implications for Genomic Selection. Anim. Int. J. Anim. Biosci. 2011, 5, 1162–1169. [Google Scholar] [CrossRef] [Green Version]
  99. Vanvanhossou, S.F.U.; Scheper, C.; Dossa, L.H.; Yin, T.; Brügemann, K.; König, S. A Multi-Breed GWAS for Morphometric Traits in Four Beninese Indigenous Cattle Breeds Reveals Loci Associated with Conformation, Carcass and Adaptive Traits. BMC Genom. 2020, 21, 783. [Google Scholar] [CrossRef] [PubMed]
  100. Igoshin, A.V.; Yurchenko, A.A.; Belonogova, N.M.; Petrovsky, D.V.; Aitnazarov, R.B.; Soloshenko, V.A.; Yudin, N.S.; Larkin, D.M. Genome-Wide Association Study and Scan for Signatures of Selection Point to Candidate Genes for Body Temperature Maintenance under the Cold Stress in Siberian Cattle Populations. BMC Genet. 2019, 20, 26. [Google Scholar] [CrossRef] [PubMed]
  101. Porto-neto, L.R.; Reverter, A.; Prayaga, K.C. The Genetic Architecture of Climatic Adaptation of Tropical Cattle. PLoS ONE 2014, 9, e113284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Raven, L.-A.; Cocks, B.G.; Hayes, B.J. Multibreed Genome Wide Association Can Improve Precision of Mapping Causative Variants Underlying Milk Production in Dairy Cattle. BMC Genomics 2014, 15, 62. [Google Scholar] [CrossRef] [Green Version]
  103. Meuwissen, T.H.; Hayes, B.J.; Goddard, M.E. Prediction of Total Genetic Value Using Genome-Wide Dense Marker Maps. Genetics 2001, 157, 1819–1829. [Google Scholar] [CrossRef]
  104. Kijas, J.W.; Townley, D.; Dalrymple, B.P.; Heaton, M.P.; Maddox, J.F.; Wilson, P.; Ingersoll, R.G.; Mcculloch, R.; Mcwilliam, S.; Tang, D.; et al. A Genome Wide Survey of SNP Variation Reveals the Genetic Structure of Sheep Breeds. PLoS ONE 2009, 4, e4668. [Google Scholar] [CrossRef] [Green Version]
  105. Michailidou, S.; Tsangaris, G.T.; Tzora, A.; Skoufos, I.; Banos, G.; Argiriou, A.; Arsenos, G. Analysis of Genome-Wide DNA Arrays Reveals the Genomic Population Structure and Diversity in Autochthonous Greek Goat Breeds. PLoS ONE 2019, 14, e0226179. [Google Scholar] [CrossRef] [Green Version]
  106. Yang, B.; Cui, L.; Enciso, M.P.; Traspov, A.; Crooijmans, R.P.M.A.; Zinovieva, N.; Schook, L.B.; Archibald, A.; Gatphayak, K.; Knorr, C. Genome-Wide SNP Data Unveils the Globalization of Domesticated Pigs. Genet. Sel. Evol. 2017, 49, 71. [Google Scholar] [CrossRef] [Green Version]
  107. Bruford, M.W.; Ginja, C.; Hoffmann, I.; Joost, S.; Orozco-terWengel, P.; Alberto, F.J.; Amaral, A.J.; Barbato, M.; Biscarini, F.; Colli, L.; et al. Prospects and Challenges for the Conservation of Farm Animal Genomic Resources, 2015–2025. Front. Genet. 2015, 6, 314. [Google Scholar] [CrossRef] [Green Version]
  108. Xu, L.; Bickhart, D.M.; Cole, J.B.; Schroeder, S.G.; Song, J.; Van Tassell, C.P.; Sonstegard, T.; Liu, G.E. Genomic Signatures Reveal New Evidences for Selection of Important Traits in Domestic Cattle. Mol. Biol. Evol. 2015, 32, 711–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Yang, I.; Kim, S. Analysis of Whole Transcriptome Sequencing Data: Workflow and Software. Genom. Inform. 2015, 13, 119–125. [Google Scholar] [CrossRef] [PubMed]
  110. Marino, R.; Capoferri, R.; Panelli, S.; Minozzi, G.; Strozzi, F.; Trevisi, E.; Snel, G.G.M.; Ajmone-Marsan, P.; Williams, J.L. Johne’s Disease in Cattle: An in Vitro Model to Study Early Response to Infection of Mycobacterium Avium Subsp Paratuberculosis Using RNA-Seq. Mol. Immunol. 2016, 91, 259–271. [Google Scholar] [CrossRef]
  111. Wang, L.; Cai, B.; Zhou, S.; Zhu, H.; Qu, L.; Wang, X.; Al, E. RNA-Seq Reveals Transcriptome Changes in Goats Following Myostatin Gene Knockout. PLoS ONE 2017, 12, e0187966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Zappaterra, M.; Gioiosa, S.; Chillemi, G.; Zambonelli, P.; Davoli, R. Muscle Transcriptome Analysis Identifies Genes Involved in Ciliogenesis and the Molecular Cascade Associated with Intramuscular Fat Content in Large White Heavy Pigs. PLoS ONE 2020, 15, e0233372. [Google Scholar] [CrossRef] [PubMed]
  113. Kern, C.; Wang, Y.; Chitwood, J. Genome-Wide Identification of Tissue-Specific Long Non-Coding RNA in Three Farm Animal Species. BMC Genom. 2018, 19, 684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Miao, X.; Luo, Q.; Zhao, H.; Al, E. Co-Expression Analysis and Identification of Fecundity-Related Long Non-Coding RNAs in Sheep Ovaries. Sci. Rep. 2016, 6, 39398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Pasquariello, R.; Fernandez-Fuertes, B.; Strozzi, F.; Mazza, R.; Lonergan, P.; Gandolfi, F.; Williams, J.L. Profiling Bovine Blastocyst MicroRNAs Using Deep Sequencing. J. Reprod. Fertil. Dev. 2017, 29, 1545–1555. [Google Scholar] [CrossRef]
  116. Low, W.Y.; Tearle, R.; Liu, R.; Koren, S.; Rhie, A.; Bickhart, D.M.; Rosen, B.D.; Kronenberg, Z.N.; Kingan, S.B.; Tseng, E.; et al. Haplotype-Resolved Genomes Provide Insights into Structural Variation and Gene Content in Angus and Brahman Cattle. Nat. Commun. 2020, 11, 2071. [Google Scholar] [CrossRef]
  117. Koren, S.; Rhie, A.; Walenz, B.P.; Dilthey, A.T.; Bickhart, D.M.; Kingan, S.B.; Hiendleder, S.; Williams, J.L.; Smith, T.P.L.; Adam, M. Complete Assembly of Parental Haplotypes with Trio Binning. bioRxiv 2018, 271486. [Google Scholar] [CrossRef] [Green Version]
  118. Jin, L.; Mao, K.; Li, J.; Huang, W.; Che, T.; Fu, Y.; Tang, Q.; Liu, P.; Song, Y.; Liu, R.; et al. Genome-Wide Profiling of Gene Expression and DNA Methylation Provides Insight into Low-Altitude Acclimation in Tibetan Pigs. Gene 2018, 5, 522–532. [Google Scholar] [CrossRef]
  119. Giuffra, E.; Tuggle, C.K.; FAANG Consortium. Functional Annotation of Animal Genomes (FAANG): Current Achievements and Roadmap. Annu. Rev. Anim. Biosci. 2018, 7, 65–88. [Google Scholar] [CrossRef]
  120. Georges, M.; Charlier, C.; Hayes, B. Harnessing Genomic Information for Livestock Improvement. Nat. Rev. Genet. 2019, 20, 135–156. [Google Scholar] [CrossRef]
  121. Cantor, R.M.; Lange, K.; Sinsheimer, J.S. Prioritizing GWAS Results: A Review of Statistical Methods and Recommendations for Their Application. Am. J. Hum. Genet. 2010, 86, 6–22. [Google Scholar] [CrossRef] [Green Version]
  122. Tam, V.; Patel, N.; Turcotte, M.; Bossé, Y.; Paré, G.; Meyre, D. Benefits and Limitations of Genome-Wide Association Studies. Nat. Rev. Genet. 2019, 20, 467–484. [Google Scholar] [CrossRef]
  123. Hatzikotoulas, K.; Gilly, A.; Zeggini, E. Using Population Isolates in Genetic Association Studies. Brief. Funct. Genom. 2014, 13, 371–377. [Google Scholar] [CrossRef] [Green Version]
  124. Sharma, A.; Seop, J.; Dang, C.G.; Sudrajad, P.; Kim, H.C. Stories and Challenges of Genome Wide Association Studies in Livestock—A Review. Asian-Australas. J. Anim. Sci. 2015, 28, 1371–1379. [Google Scholar] [CrossRef] [PubMed]
  125. Hayes, B.J.; Lewin, H.A.; Goddard, M.E. The Future of Livestock Breeding: Genomic Selection for Efficiency, Reduced Emissions Intensity, and Adaptation. Trends Genet. 2013, 29, 206–214. [Google Scholar] [CrossRef]
  126. Hayes, B. Overview of Statistical Methods for Genome-Wide Association Studies (GWAS). Methods Mol. Biol. 2013, 1019, 149–169. [Google Scholar] [PubMed]
  127. Goddard, M.E.; Hayes, B.J. Mapping Genes for Complex Traits in Domestic Animals and Their Use in Breeding Programmes. Nat. Rev. 2009, 10, 381–391. [Google Scholar] [CrossRef]
  128. Bouwman, A.C.; Daetwyler, H.D.; Chamberlain, A.J.; Ponce, C.H.; Sargolzaei, M.; Schenkel, F.S.; Sahana, G.; Govignon-gion, A.; Boitard, S.; Dolezal, M.; et al. Meta-Analysis of Genome-Wide Association Studies for Cattle Stature Identifies Common Genes That Regulate Body Size in Mammals. Nat. Genet. 2018, 50, 362–367. [Google Scholar] [CrossRef]
  129. Kominakis, A.; Hager-Theodorides, A.L.; Zoidis, E.; Saridaki, A.; Antonakos, G.; Tsiamis, G. Combined GWAS and ‘Guilt by Association’-Based Prioritization Analysis Identifies Functional Candidate Genes for Body Size in Sheep. Genet. Sel. Evol. 2017, 49, 41. [Google Scholar] [CrossRef]
  130. Braz, C.U.; Rowan, T.N.; Schnabel, R.D.; Decker, J.E. Genome-wide Association Analyses Identify Genotype-by-environment Interactions of Growth Traits in Simmental Cattle. Sci. Rep. 2021, 11, 13335. [Google Scholar] [CrossRef] [PubMed]
  131. Koch, R.M.; Swiger, L.; Chambers, D.; Gregory, K. Efficiency of Feed Use in Beef Cattle. J. Anim. Sci. 1963, 22, 486–494. [Google Scholar] [CrossRef]
  132. Herd, R.M.; Bishop, S.C. Genetic Variation in Residual Feed Intake and Its Association with Other Production Traits in British Hereford Cattle. Livest. Prod. Sci. 2000, 63, 111–119. [Google Scholar] [CrossRef]
  133. Tortereau, F.; Weisbecker, J.; Marcon, D.; Bouvier, F.; François, D. Genetic Parameters for Feed Efficiency in Romane Rams and Responses to Single-Generation Selection. Anim. Int. J. Anim. Biosci. 2020, 14, 681–687. [Google Scholar] [CrossRef]
  134. Santana, M.H.A.; Utsunomiya, Y.T.; Neves, H.H.R.; Gomes, R.C.; Garcia, J.F.; Fukumasu, H.; Silva, S.L.; Oliveira Junior, G.A.; Alexandre, P.A.; Leme, P.R.; et al. Genome-Wide Association Analysis of Feed Intake and Residual Feed Intake in Nellore Cattle. BMC Genet. 2014, 15, 21. [Google Scholar] [CrossRef] [Green Version]
  135. De Oliveira, P.S.N.; Cesar, A.S.M.; Nascimento, M.L.; Chaves, A.S.; Tizioto, P.C.; Tullio, R.R.; Lanna, D.P.D.; Rosa, A.N.; Sonstegard, T.S.; Mourao, G.B.; et al. Identification of Genomic Regions Associated with Feed Efficiency in Nelore Cattle. BMC Genet. 2014, 15, 100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Riley, D.G.; Gill, C.A.; Boldt, C.R.; Funkhouser, R.R.; Herring, A.D.; Riggs, P.K.; Sawyer, J.E.; Lunt, D.K.; Sanders, J.O. Crossbred Steer Temperament as Yearlings and Whole Genome Association of Steer Temperament as Yearlings and Calf Temperament Post-Weaning. J. Anim. Sci. 2016, 94, 1408–1414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Vallée, A.; Daures, J.; van Arendonk, J.A.M.; Bovenhuis, H. Genome-Wide Association Study for Behavior, Type Traits, and Muscular Development in Charolais Beef Cattle. J. Anim. Sci. 2016, 94, 2307–2316. [Google Scholar] [CrossRef]
  138. Paredes-Sanchez, F.A.; Sifuentes-Rincon, A.M.; Casas, E.; Arellano-vera, W.; Parra-bracamonte, G.M.; Riley, D.G.; Welsh, T.H., Jr.; Randel, R.D. Novel Genes Involved in the Genetic Architecture of Temperament in Brahman Cattle. PLoS ONE 2020, 15, e0237825. [Google Scholar] [CrossRef]
  139. Vitti, J.J.; Grossman, S.R.; Sabeti, P.C. Detecting Natural Selection in Genomic Data. Annu. Rev. Genet. 2013, 47, 97–120. [Google Scholar] [CrossRef] [PubMed]
  140. Zeng, J.; De Vlaming, R.; Wu, Y.; Robinson, M.R.; Lloyd-Jones, L.R.; Yengo, L.; Yap, C.X.; Xue, A.; Sidorenko, J.; McRae, A.F.; et al. Signatures of Negative Selection in the Genetic Architecture of Human Complex Traits. Nat. Genet. 2018, 50, 746–753. [Google Scholar] [CrossRef]
  141. Charlesworth, D. Balancing Selection and Its Effects on Sequences in Nearby Genome Regions. PLoS Genet. 2006, 2, 379–384. [Google Scholar] [CrossRef] [Green Version]
  142. Barton, N.H. Genetic Hitchhiking. Trans. R. Soc. B 2000, 355, 553–1562. [Google Scholar] [CrossRef] [PubMed]
  143. Cao, Y.H.; Xu, S.S.; Shen, M.; Chen, Z.H.; Gao, L.; Lv, F.H.; Xie, X.L.; Wang, X.H.; Yang, H.; Liu, C.-B.; et al. Historical Introgression from Wild Relatives Enhanced Climatic Adaptation and Resistance to Pneumonia in Sheep. Mol. Biol. Evol. 2021, 38, 838–855. [Google Scholar] [CrossRef] [PubMed]
  144. Yurchenko, A.A.; Daetwyler, H.D.; Yudin, N.; Schnabel, R.D.; Vander Jagt, C.J.; Soloshenko, V.; Lhasaranov, B.; Popov, R.; Taylor, J.F.; Larkin, D.M. Scans for Signatures of Selection in Russian Cattle Breed Genomes Reveal New Candidate Genes for Environmental Adaptation and Acclimation. Sci. Rep. 2018, 8, 12984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Chen, Y.; Yang, Y.; Li, C.; Li, R.; Xiao, H.; Chen, S. Genetic Diversity of TLR3 and TLR8 Genes among Five Chinese Native Cattle Breeds from Southwest China. Livest. Sci. 2020, 232, 103895. [Google Scholar] [CrossRef]
  146. Eusebi, P.G.; Cortés, O.; Carleos, C.; Dunner, S.; Cañon, J. Detection of Selection Signatures for Agonistic Behaviour in Cattle. J. Anim. Breed. Genet. 2018, 135, 170–177. [Google Scholar] [CrossRef]
  147. Eusebi, P.G.; Sevane, N.; Cortés, O.; Contreras, E.; Cañon, J.; Dunner, S. Aggressive Behavior in Cattle Is Associated with a Polymorphism in the MAOA Gene Promoter. Anim. Genet. 2020, 51, 14–21. [Google Scholar] [CrossRef]
  148. Hamlyn-Hill, F. Improving Temperament: Effects on Productivity and Meat Quality. Beef CRC. Available online: http://futurebeef.com.au/topics/breeding-and-genetics/improving-temperament-andflight-time (accessed on 28 July 2021).
  149. Burdick, N.; Agado, B.; White, J.; Matheney, K.; Neuendorff, D.; Riley, D.; Vann, R.C.; Welsh, T.H., Jr.; Randel, R.D. Technical Note: Evolution of Exit Velocity in Suckling Brahman Calves. J. Anim. Sci. 2011, 89, 233–236. [Google Scholar] [CrossRef] [Green Version]
  150. Café, L.M.; Robinson, D.L.; Ferguson, D.M.; McIntyre, B.L.; Geesink, G.H.; Greenwood, P.L. Cattle Temperament: Persistence of Assessments and Associations with Productivity, Efficiency, Carcass and Meat Quality Traits. J. Anim. Sci. 2011, 89, 1452–1465. [Google Scholar] [CrossRef]
  151. Burrow, H. Variances and Covariances between Productive and Adaptative Traits and Temperament in a Composite Breed of Tropical Beef Cattle. Livest. Prod. Sci. 2001, 70, 213–233. [Google Scholar] [CrossRef]
  152. Hoppe, S.; Brandt, H.R.; Nig, S.K.; Erhardt, G.; Gauly, M. Temperament Traits of Beef Calves Measured under Field Condi-Tions and Their Relationships to Performance. J. Anim. Sci. 2010, 88, 1892–1898. [Google Scholar] [CrossRef]
  153. Lv, F.; Agha, S.; Kantanen, J.; Colli, L.; Stucki, S.; Kijas, J.W.; Li, M.; Marsan, P.A. Adaptations to Climate-Mediated Selective Pressures in Sheep. Mol. Biol. Evol. 2014, 31, 3324–3343. [Google Scholar] [CrossRef] [Green Version]
  154. Igoshin, A.; Yudin, N.; Aitnazarov, R.; Yurchenko, A.A.; Larkin, D.M. Whole-Genome Resequencing Points to Candidate DNA Loci Affecting Body Temperature under Cold Stress in Siberian Cattle Populations. Life 2021, 11, 959. [Google Scholar] [CrossRef]
  155. Li, R.; Li, H.; Chen, R.; Chong, Q.; Xiao, H.; Chen, S. Genome-Wide Scan of Selection Signatures in Dehong Humped Cattle for Heat Tolerance and Disease Resistance. Anim. Genet. 2020, 51, 292–299. [Google Scholar]
  156. Edea, Z.; Dadi, H.; Dessie, T.; Kim, K.S. Genomic Signatures of High-Altitude Adaptation in Ethiopian Sheep Popula-Tions. Genes Genom. 2019, 41, 973–981. [Google Scholar] [CrossRef]
  157. Freitas, P.H.F.; Wang, Y.; Yan, P.; Oliveira, H.R.; Schenkel, F.S.; Zhang, Y.; Brito, L.F. Genetic Diversity and Signatures of Selection for Thermal Stress in Cattle and Other Two Bos Species Adapted to Divergent Climatic Conditions. Front. Genet. 2021, 12, 102. [Google Scholar] [CrossRef]
  158. Buggiotti, L.; Yurchenko, A.A.; Yudin, N.S.; Vander Jagt, C.J.; Nadezhda, V.; Kusliy, M.; Vasiliev, S.K.; Rodionov, A.N.; Boronetskaya, O.I.; Zinovieva, A.; et al. Demographic History, Adaptation, and NRAP Convergent Evolution at Amino Acid Residue 100 in the World Northernmost Cattle from Siberia. Mol. Biol. Evol. 2021, 38, 3093–3110. [Google Scholar] [CrossRef]
  159. Wang, X.; Ju, Z.; Jiang, Q.; Zhong, J.; Liu, C.; Wang, J.; Hoff, J.L.; Schnabel, R.D.; Zhao, H.; Gao, Y.; et al. Introgression, Admixture and Selection Facilitate Genetic Adaptation to High-Altitude Environments in Cattle. Genomics 2021, 113, 1491–1503. [Google Scholar] [CrossRef]
  160. Molotsi, A.; Cloete, S.; Taylor, J.F.; Whitacre, L. Identification of Selection Signatures in South African Sheep Popula-Tions Using HAPFLK and Bayesian Fst Approaches. In Proceedings of the World Congress on Genetics Applied to Livestock, Auckland, New Zealand, 7–11 February 2018; pp. 1–5. [Google Scholar]
  161. Mwacharo, J.M.; Kim, E.S.; Elbeltagy, A.R.; Aboul-Naga, A.M.; Rischkowsky, B.A.; Rothschild, M.F. Genomic Footprints of Dryland Stress Adaptation in Egyptian Fat-Tail Sheep and Their Divergence from East African and Western Asia Cohorts. Sci. Rep. 2017, 7, 17647. [Google Scholar] [CrossRef] [Green Version]
  162. Saravanan, K.A.; Panigrahi, M.; Kumar, H.; Bhushan, B.; Dutt, T.; Mishra, B.P. Genome-Wide Analysis of Genetic Diversity and Selection Signatures in Three Indian Sheep Breeds. Livest. Sci. 2021, 243, 104367. [Google Scholar] [CrossRef]
  163. Álvarez, I.; Fernández, I.; Traoré, A.; Pérez-Pardal, L.; Menéndez-Arias, N.A.; Goyache, F. Ancient Homozygosity Segments in West African Djallonké Sheep Inform on the Genomic Impact of Livestock Adaptation to the Environment. Animals 2020, 10, 1178. [Google Scholar] [CrossRef]
  164. Bertolini, F.; Servin, B.; Talenti, A.; Rochat, E.; Kim, E.S.; Oget, C.; Palhière, I.; Crisà, A.; Catillo, G.; Steri, R.; et al. Signatures of Selection and Environmental Adaptation across the Goat Genome Post-Domestication. Genet. Sel. Evol. 2018, 50, 57. [Google Scholar] [CrossRef]
  165. Kim, E.S.; Elbeltagy, A.R.; Aboul-Naga, A.M.; Rischkowsky, B.; Sayre, B.; Mwacharo, J.M.; Rothschild, M.F. Multiple Genomic Signatures of Selection in Goats and Sheep Indigenous to a Hot Arid Environment. Heredity (Edinb) 2016, 116, 255–264. [Google Scholar] [CrossRef]
  166. Pitt, D.; Bruford, M.W.; Barbato, M.; Orozco-terWengel, P.; Martínez, R.; Sevane, N. Demography and Rapid Local Adaptation Shape Creole Cattle Genome Diversity in the Tropics. Evol. Appl. 2019, 12, 105–122. [Google Scholar] [CrossRef]
  167. Dutta, P.; Talenti, A.; Young, R.; Jayaraman, S.; Callaby, R.; Jadhav, S.K.; Dhanikachalam, V.; Manikandan, M.; Biswa, B.B.; Low, W.Y.; et al. Whole Genome Analysis of Water Buffalo and Global Cattle Breeds Highlights Convergent Signatures of Domestication. Nat. Commun. 2020, 11. [Google Scholar] [CrossRef]
  168. Singh, A.; Mehrotra, A.; Gondro, C.; da Silva Romero, A.R.; Pandey, A.K.; Karthikeyan, A.; Bashir, A.; Mishra, B.P.; Dutt, T.; Kumar, A. Signatures of Selection in Composite Vrindavani Cattle of India. Front. Genet. 2020, 11, 589496. [Google Scholar] [CrossRef] [PubMed]
  169. Zhang, Y.; Xue, X.; Liu, Y.; Abied, A.; Ding, Y.; Zhao, S.; Wang, W.; Ma, L.; Guo, J.; Guan, W.; et al. Genome-Wide Comparative Analyses Reveal Selection Signatures Underlying Adaptation and Production in Tibetan and Poll Dorset Sheep. Sci. Rep. 2021, 11, 2466. [Google Scholar] [CrossRef]
  170. Álvarez, I.; Fernández, I.; Traoré, A.; Pérez-Pardal, L.; Menéndez-Arias, N.A.; Goyache, F. Genomic Scan of Selective Sweeps in Djallonké (West African Dwarf) Sheep Shed Light on Adaptation to Harsh Environments. Sci. Rep. 2020, 10, 2824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Eydivandi, S.; Roudbar, M.A.; Ardestani, S.S.; Momen, M.; Sahana, G. A Selection Signatures Study among Middle Eastern and European Sheep Breeds. J. Anim. Breed. Genet. 2021, 138, 1–15. [Google Scholar] [CrossRef] [PubMed]
  172. Fan, R.; Gu, Z.; Guang, X.; Marín, J.C.; Varas, V.; González, B.A.; Wheeler, J.C.; Hu, Y.; Li, E.; Sun, X.; et al. Genomic Analysis of the Domestication and Post-Spanish Conquest Evolution of the Llama and Alpaca. Genome Biol. 2020, 21, 159. [Google Scholar] [CrossRef]
  173. Macciotta, N.P.P.; Colli, L.; Cesarani, A.; Ajmone-Marsan, P.; Low, W.Y.; Tearle, R.; Williams, J.L. The Distribution of Runs of Homozygosity in the Genome of River and Swamp Buffaloes Reveals a History of Adaptation, Migration and Crossbred Events. Genet. Sel. Evol. 2021, 53, 20. [Google Scholar] [CrossRef]
  174. Abied, A.; Xu, L.; Sahlu, B.W.; Xing, F.; Ahbara, A.; Pu, Y.; Lin, J.; Berihulay, H.; Islam, R.; He, X.; et al. Genome-Wide Analysis Revealed Homozygosity and Demographic History of Five Chinese Sheep Breeds Adapted to Different Environments. Genes 2020, 11, 1480. [Google Scholar] [CrossRef]
  175. Tajima, F. Statistical Method for Testing the Neutral Mutation Hypothesis by DNA Polymorphism. Genetics 1989, 595, 585–595. [Google Scholar] [CrossRef]
  176. Simonsen, K.L.; Churchill, G.A.; Aquadro, C.F. Properties of Statistical Tests of Neutrality for DNA Polymorphism Data. Genetics 1995, 141, 413–429. [Google Scholar] [CrossRef]
  177. Fay, J.C.; Wu, C.I. Hitchhiking under Positive Darwinian Selection. Genetics 2000, 155, 1405–1413. [Google Scholar] [CrossRef] [PubMed]
  178. Saravanan, K.A.; Panigrahi, M.; Kumar, H.; Bhushan, B.; Dutt, T.; Mishra, B.P. Selection Signatures in Livestock Genome: A Review of Concepts, Approaches and Applications. Livest. Sci. 2020, 241, 104257. [Google Scholar] [CrossRef]
  179. Fariello, M.I.; Boitard, S.; Naya, H.; SanCristobal, M.; Servin, B. Detecting Signatures of Selection through Haplotype Differentiation among Hierarchically Structured Populations. Genetics 2013, 193, 929–941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Kim, Y.; Neilsen, R. Linkage Disequilibrium as a Signature of Selective Sweeps. Genetics 2004, 167, 1513–1524. [Google Scholar] [CrossRef] [Green Version]
  181. Sabeti, P.C.; Reich, D.E.; Higgins, J.M.; Levine, H.Z.P.; Richter, D.J.; Schaffner, S.F.; Gabriel, S.B.; Platko, J.V.; Patterson, N.J.; McDonald, G.J.; et al. Detecting Recent Positive Selection in the Human Genome from Haplotype Structure. Nature 2002, 419, 832–837. [Google Scholar] [CrossRef]
  182. Sabeti, P.C.; Varilly, P.; Fry, B.; Lohmueller, J.; Hostetter, E.; Cotsapas, C.; Xie, X.; Byrne, E.H.; McCarroll, S.A.; Gaudet, R.; et al. Genome-Wide Detection and Characterization of Positive Selection in Human Populations. Nature 2007, 449, 913–918. [Google Scholar] [CrossRef]
  183. Voight, B.F.; Kudaravalli, S.; Wen, X.; Pritchard, J.K. A Map of Recent Positive Selection in the Human Genome. PLoS Biol. 2006, 4, e72. [Google Scholar] [CrossRef] [Green Version]
  184. Tang, H.; Choudhry, S.; Mei, R.; Morgan, M.; Rodriguez-Cintron, W.; Burchard, E.G.; Risch, N.J. Recent Genetic Selection in the Ancestral Admixture of Puerto Ricans. Am. J. Hum. Genet. 2007, 81, 626–633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Gibson, J.; Morton, N.E.; Collins, A. Extended Tracts of Homozygosity in Outbred Human Populations. Hum. Mol. Genet. 2006, 15, 789–795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Brisbin, A.; Bryc, K.; Byrnes, J.; Zakharia, F.; Omberg, L.; Degenhardt, J.; Reynolds, A.; Ostrer, H.; Mezey, J.G.; Bustamante, C.D. PCAdmix: Principal Components-Based Assignment of Ancestry along Each Chromosome in Individuals with Admixed Ancestry from Two or More Populations. Hum. Biol. 2012, 84, 343–364. [Google Scholar] [CrossRef] [Green Version]
  187. Atkinson, E.G.; Maihofer, A.X.; Kanai, M.; Martin, A.R.; Karczewski, K.J.; Santoro, M.L.; Ulirsch, J.C.; Kamatani, Y.; Okada, Y.; Finucane, H.K.; et al. Tractor Uses Local Ancestry to Enable the Inclusion of Admixed Individuals in GWAS and to Boost Power. Nat. Genet. 2021, 53, 195–204. [Google Scholar] [CrossRef]
  188. Pasaniuc, B.; Sankararaman, S.; Kimmel, G.; Halperin, E. Inference of Locus-Specific Ancestry in Closely Related Populations. Bioinformatics 2009, 25, i213–i221. [Google Scholar] [CrossRef] [Green Version]
  189. Schubert, R.; Andaleon, A.; Wheeler, H.E. Comparing Local Ancestry Inference Models in Populations of Two- And Three-Way Admixture. PeerJ 2020, 8, e10090. [Google Scholar] [CrossRef]
  190. Wu, J.; Liu, Y.; Zhao, Y. Systematic Review on Local Ancestor Inference From a Mathematical and Algorithmic Perspective. Front. Genet. 2021, 12, 639877. [Google Scholar] [CrossRef]
  191. Tang, H.; Coram, M.; Wang, P.; Zhu, X.; Risch, N. Reconstructing Genetic Ancestry Blocks in Admixed Individuals. Am. J. Hum. Genet. 2006, 79, 1–12. [Google Scholar] [CrossRef] [Green Version]
  192. Guan, Y. Detecting Structure of Haplotypes and Local Ancestry. Genetics 2014, 196, 625–642. [Google Scholar] [CrossRef] [Green Version]
  193. Salter-Townshend, M.; Myers, S. Fine-Scale Inference of Ancestry Segments without Prior Knowledge of Admixing Groups. Genetics 2019, 212, 869–889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Maples, B.K.; Gravel, S.; Kenny, E.E.; Bustamante, C.D. RFMix: A Discriminative Modeling Approach for Rapid and Robust Local-Ancestry Inference. Am. J. Hum. Genet. 2013, 93, 278–288. [Google Scholar] [CrossRef] [Green Version]
  195. Dias-Alves, T.; Mairal, J.; Blum, M.G.B. Loter: A Software Package to Infer Local Ancestry for a Wide Range of Species. Mol. Biol. Evol. 2018, 35, 2318–2326. [Google Scholar] [CrossRef] [Green Version]
  196. Utsunomiya, Y.T.; Milanesi, M.; Barbato, M.; Utsunomiya, A.T.H.; Sölkner, J.; Ajmone-Marsan, P.; Garcia, J.F. Unsupervised Detection of Ancestry Tracks with the GHap R Package. Methods Ecol. Evol. 2020, 11, 1448–1454. [Google Scholar] [CrossRef]
  197. Popescu, A.-A.; Huber, K.T. PSIKO2: A Fast and Versatile Tool to Infer Population Stratification on Various Levels in GWAS. Bioinformatics 2015, 31, 3552–3554. [Google Scholar] [CrossRef] [Green Version]
  198. Zhao, W.; Ma, Y.; Chen, A.; Fu, L.; Zhang, Y. Robust Sliding Mode Control for Markovian Jump Singular Systems with Randomly Changing Structure. Appl. Math. Comput. 2019, 349, 81–96. [Google Scholar] [CrossRef]
  199. Barbato, M.; Hailer, F.; Orozco-terWengel, P.; Kijas, J.W.; Mereu, P.; Cabras, P.; Mazza, R.; Pirastru, M.; Bruford, M.W. Genomic Signatures of Adaptive Introgression from European Mouflon into Domestic Sheep. Sci. Rep. 2017, 7, 7623. [Google Scholar] [CrossRef] [Green Version]
  200. Barbato, M.; Hailer, F.; Upadhyay, M.; Del Corvo, M.; Colli, L.; Negrini, R.; Kim, E.; Crooijmans, R.P.M.A.; Sonstegard, T.; Ajmone-marsan, P. Adaptive Introgression from Indicine Cattle into White Cattle Breeds from Central Italy. Sci. Rep. 2020, 10, 1279. [Google Scholar] [CrossRef]
  201. Wu, D.-D.; Ding, X.-D.; Wang, S.; Wójcik, J.M.; Zhang, Y.; Tokarska, M.; Li, Y.; Wang, M.-S.; Faruque, O.; Nielsen, R.; et al. Pervasive Introgression Facilitated Domestication and Adaptation in the Bos Species Complex. Nat. Ecol. Evol. 2018, 2, 1139–1145. [Google Scholar] [CrossRef]
  202. Hu, X.-J.; Yang, J.; Xie, X.-L.; Lv, F.-H.; Cao, Y.-H.; Li, W.-R.; Liu, M.-J.; Wang, Y.-T.; Li, J.-Q.; Liu, Y.-G.; et al. The Genome Landscape of Tibetan Sheep Reveals Adaptive Introgression from Argali and the History of Early Human Settlements on the Qinghai–Tibetan Plateau. Mol. Biol. Evol. 2019, 36, 283–303. [Google Scholar] [CrossRef] [Green Version]
  203. Kim, K.; Kwon, T.; Dessie, T.; Yoo, D.; Mwai, O.A.; Jang, J.; Sung, S.; Lee, S.; Salim, B.; Jung, J.; et al. The Mosaic Genome of Indigenous African Cattle as a Unique Genetic Resource for African Pastoralism. Nat. Genet. 2020, 52, 1099–1110. [Google Scholar] [CrossRef]
  204. Manel, S.; Schwartz, M.K.; Luikart, G.; Taberlet, P. Landscape Genetics: Combining Landscape Ecology and Population Genetics. Trends Ecol. Evol. 2003, 18, 189–197. [Google Scholar] [CrossRef]
  205. Jelinski, D.E. On Genes and Geography: A Landscape Perspective on Genetic Variation in Natural Plant Populations. Landsc. Urban Plan. 1997, 39, 11–23. [Google Scholar] [CrossRef]
  206. Goodchild, M. Geographical Information Science. Int. J. Geogr. Inf. Syst. 1992, 6, 31–45. [Google Scholar] [CrossRef]
  207. Joost, S.; Bonin, A.; Bruford, M.; Despres, L.; Concord, C.; Erhardt, G.; Taberlet, P. A Spatial Analysis Method (SAM) to Detect Candidate Loci for Selection: Towards a Landscape Genomics Approach to Adaptation. Mol. Ecol. 2007, 16, 3955–3969. [Google Scholar] [CrossRef]
  208. Joost, S.; Kalbermatten, M.; Bonin, A. Spatial Analysis Method (Sam): A Software Tool Combining Molecular and Environmental Data to Identify Candidate Loci for Selection. Mol. Ecol. Resour. 2008, 8, 957–960. [Google Scholar] [CrossRef]
  209. Pariset, L.; Joost, S.; Ajmone Marsan, P.; Valentini, A.; Econogene Consortium (EC). Landscape Genomics and Biased F ST Approaches Reveal Single Nucleotide Polymorphisms under Selection in Goat Breeds of North-East Mediterranean. BMC Genet. 2009, 10, 7. [Google Scholar] [CrossRef] [Green Version]
  210. Tonteri, A.; Vasemagi, A.; Lumme, J.; Primmer, C.R. Beyond MHC: Signals of Elevated Selection Pressure on Atlantic Salmon (Salmo salar) Immune-Relevant Loci. Mol. Ecol. 2010, 19, 1273–1282. [Google Scholar] [CrossRef]
  211. Frichot, E.; Schoville, S.D.; Bouchard, G.; Franc, O. Testing for Associations between Loci and Environmental Gradients Using Latent Factor Mixed Models. Mol. Biol. Evol. 2013, 30, 1687–1699. [Google Scholar] [CrossRef] [Green Version]
  212. Mdladla, K.; Dzomba, E.F. Landscape Genomics and Pathway Analysis to Understand Genetic Adaptation of South African Indigenous Goat Populations. Heredity (Edinb) 2018, 120, 369–378. [Google Scholar] [CrossRef]
  213. Stucki, S.; Orozco-Terwengel, P.; Forester, B.R.; Duruz, S.; Colli, L.; Masembe, C. High Performance Computation of Landscape Genomic Models Including Local Indicators of Spatial Association. Mol. Ecol. Resour. 2017, 17, 1072–1089. [Google Scholar] [CrossRef] [Green Version]
  214. Cortellari, M.; Barbato, M.; Talenti, A.; Bionda, A.; Randi, E.; Sarti, F.M.; Sartore, S.; Soglia, D.; Liotta, L. The Climatic and Genetic Heritage of Italian Goat Breeds with Genomic SNP Data. Sci. Rep. 2021, 11, 10986. [Google Scholar] [CrossRef]
  215. Wang, Z.; Zhang, H.; Yang, H.; Wang, S.; Rong, E.; Pei, W.; Li, H.; Wang, N. Genome-Wide Association Study for Wool Production Traits in a Chinese Merino Sheep Population. PLoS ONE 2014, 9, e107101. [Google Scholar] [CrossRef] [PubMed]
  216. Silva, D.B.S.; Fonseca, L.F.S.; Pinheiro, D.G.; Magalhães, A.F.B.; Muniz, M.M.M.; Ferro, J.A.; Baldi, F.; Chardulo, L.A.L.; Schnabel, R.D.; Taylor, J.F.; et al. Spliced Genes in Muscle from Nelore Cattle and Their Association with Carcass and Meat Quality. Sci. Rep. 2020, 10, 14701. [Google Scholar] [CrossRef]
  217. Colli, L.; Negrini, R.; Nicoloso, L.; Crepaldi, P. Assessing The Spatial Dependence of Adaptive Loci in 43 European and Western Asian Goat Breeds Using AFLP Markers. PLoS ONE 2014, 9, e86668. [Google Scholar] [CrossRef] [PubMed]
  218. Buitkamp, J.; Filmether, P.; Stear, M.; Epplen, J. Class I and Class II Major Histocompatibility Complex Alleles Are Associated with Faecal Egg Counts Following Natural, Predominantly Ostertagia Circumcincta Infection. Parasitol. Res. 1996, 82, 693–696. [Google Scholar] [CrossRef] [PubMed]
  219. Duruz, S.; Rochat, E.; Selmoni, O.; Vajana, E.; Orozco, P.; Joost, S.; Dunner, S.; Bruford, C.M.W. Rapid Identification and Interpretation of Gene—Environment Associations Using the New R. SamBada Landscape Genomics Pipeline. Mol. Ecol. Resour. 2019, 19, 1355–1365. [Google Scholar] [CrossRef] [Green Version]
  220. Gunther, T.; Coop, G. Robust Identification of Local Adaptation from Allele Frequencies. Genetics 2013, 195, 205–220. [Google Scholar] [CrossRef] [Green Version]
  221. Frichot, E.; Francois, O. LEA: An R Package for Landscape and Ecological Association Studies. Methods Ecol. Evol. 2015, 6, 925–929. [Google Scholar] [CrossRef]
  222. Guillot, G.; Vitalis, R. Detecting Correlation between Allele Frequencies and Environmental Variables as a Signature of Selection. A Fast Computational Approach for Genome-Wide Studies. Spat. Stat. 2014, 8, 145–155. [Google Scholar] [CrossRef] [Green Version]
  223. Coop, G.; Witonsky, D.; Di Rienzo, A.; Pritchard, J.K. Using Environmental Correlations to Identify Loci Underlying Local Adaptation. Genetics 2010, 185, 1411–1423. [Google Scholar] [CrossRef] [Green Version]
  224. Gautier, M. Genome-Wide Scan for Adaptive Divergence and Association with Population-Specific Covariates. Genetics 2015, 201, 1555–1579. [Google Scholar] [CrossRef] [Green Version]
  225. De Villemereuil, P.; Gaggiotti, O.E. A New F ST-Based Method to Uncover Local Adaptation Using Environmental Variables. Methods Ecol. Evol. 2015, 6, 1248–1258. [Google Scholar] [CrossRef] [Green Version]
  226. Caye, K.; Jumentier, B.; Lepeule, J.; François, O. LFMM 2: Fast and Accurate Inference of Gene-Environment Associations in Genome-Wide Studies. Mol. Biol. Evol. 2019, 36, 852–860. [Google Scholar] [CrossRef]
  227. Antao, T.; Lopes, A.; Lopes, R.J.; Beja-pereira, A.; Luikart, G. LOSITAN: A Workbench to Detect Molecular Adaptation Based on a F St -Outlier Method. BMC Bioinform. 2008, 5, 323. [Google Scholar] [CrossRef] [Green Version]
  228. Beaumont, M.A.; Nichols, R.A. Evaluating Loci for Use in the Genetic Analysis of Population Structure. Proc. R. Soc. Lond. B. 1996, 263, 1619–1623. [Google Scholar]
  229. Foll, M.; Gaggiotti, O. A Genome-Scan Method to Identify Selected Loci Appropriate for Both Dominant and Codominant Markers: A Bayesian Perspective. Genetics 2008, 993, 977–993. [Google Scholar] [CrossRef] [Green Version]
  230. Landguth, E.; Cushman, S.; Schwartz, M.; McKelvey, K.; Murphy, M.; Luikart, G. Quantifying the Lag Time to Detect Barriers in Landscape Genetics. Mol. Ecol. 2010, 19, 4179–4191. [Google Scholar] [CrossRef] [PubMed]
  231. De Mita, S.; Thuillet, A.; Gay, L.; Ahmadi, N.; Manel, S.; Ronfort, J. Detecting Selection along Environmental Gradients: Analysis of Eight Methods and Their Effectiveness for Outbreeding and Selfing Populations. Mol. Ecol. 2013, 22, 1383–1399. [Google Scholar] [CrossRef] [PubMed]
  232. Terasaki Hart, D.E.; Bishop, A.P.; Wang, I.J. Geonomics: Forward-Time, Spatially Explicit, and Arbitrarily Complex Landscape Genomic Simulations. Mol. Biol. Evol. 2021. [Google Scholar] [CrossRef]
  233. Stephens, Z.D.; Lee, S.Y.; Faghri, F.; Campbell, R.H.; Zhai, C.; Efron, M.J.; Iyer, R.; Schatz, M.C.; Sinha, S.; Robinson, G.E. Big Data: Astronomical or Genomical? PLoS Biol. 2015, 13, e1002195. [Google Scholar] [CrossRef]
  234. Neethirajan, S. Sensing and Bio-Sensing Research The Role of Sensors, Big Data and Machine Learning in Modern Animal Farming. Sens. Bio-Sens. Res. 2020, 29, 100367. [Google Scholar] [CrossRef]
  235. Rees, J.S.; Castellano, S.; Andrés, A.M. The Genomics of Human Local Adaptation. Trends Genet. 2020, 36, 415–428. [Google Scholar] [CrossRef]
  236. Schrider, D.R.; Kern, A.D. Soft Sweeps Are the Dominant Mode of Adaptation in the Human Genome. Mol. Biol. Evol. 2017, 34, 1863–1877. [Google Scholar] [CrossRef] [Green Version]
  237. Sugden, L.A.; Atkinson, E.G.; Fischer, A.P.; Rong, S.; Henn, B.M.; Ramachandran, S. Localization of Adaptive Variants in Human Genomes Using Averaged One-Dependence Estimation. Nat. Commun. 2018, 9, 703. [Google Scholar] [CrossRef] [Green Version]
  238. Nayeri, S.; Sargolzaei, M.; Tulpan, D. A Review of Traditional and Machine Learning Methods Applied to Animal Breeding. Anim. Health Res. Rev. 2019, 20, 31–46. [Google Scholar] [CrossRef]
  239. Okser, S.; Pahikkala, T.; Airola, A.; Salakoski, T.; Ripatti, S.; Aittokallio, T. Regularized Machine Learning in the Genetic Prediction of Complex Traits. PLoS Genet. 2014, 10, e1004754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Arpanahi Abdollahi, R.; Gianola, D.; Peñagaricano, F. Deep Learning versus Parametric and Ensemble Methods for Genomic Prediction of Complex Phenotypes. Genet. Sel. Evol. 2020, 52, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Li, B.; Zhang, N.; Wang, Y.-G.; George, A.W.; Reverter, A.; Li, Y. Genomic Prediction of Breeding Values Using a Subset of SNPs Identified by Three Machine Learning Methods. Front. Genet. 2018, 9, 237. [Google Scholar] [CrossRef]
  242. Waldmann, P. Approximate Bayesian Neural Networks in Genomic Prediction. Genet. Sel. Evol. 2018, 50, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Bellot, P.; Campos, G.D.L.; Pérez-enciso, M. Can Deep Learning Improve Genomic Prediction of Complex Human Traits? Genetics 2018, 210, 809–819. [Google Scholar] [CrossRef] [Green Version]
  244. Piles, M.; Lozano, C.F.; Galilea, M.V.; Rodríguez, O.G.; Sánchez, J.P.; Torrallardona, D.; Ballester, M.; Quintanilla, R. Machine Learning Applied to Transcriptomic Data to Identify Genes Associated with Feed Efficiency in Pigs. Genet. Sel. Evol. 2019, 51, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Kaluarachchi, T.; Reis, A.; Nanayakkara, S. A Review of Recent Deep Learning Approaches in Human-Centered Machine Learning. Sensors 2021, 21, 2514. [Google Scholar] [CrossRef] [PubMed]
  246. Helm, J.M.; Swiergosz, A.M.; Haeberle, H.S.; Karnuta, J.M.; Schaffer, J.L.; Krebs, V.E.; Spitzer, A.I.; Ramkumar, P.N. Machine Learning and Artificial Intelligence: Definitions, Applications, and Future Directions. Curr. Rev. Musculoskelet. Med. 2020, 13, 69–76. [Google Scholar] [CrossRef] [PubMed]
  247. Jordan, M.I.; Mitchell, T.M. Machine Learning: Trends, Perspectives, and Prospects. Science 2015, 349, 255–260. [Google Scholar] [CrossRef] [PubMed]
  248. Schrider, D.R.; Kern, A.D. Supervised Machine Learning for Population Genetics: A New Paradigm. Trends Genet. 2018, 34, 301–312. [Google Scholar] [CrossRef] [Green Version]
  249. Harfouche, A.L.; Jacobson, D.A.; Kainer, D.; Romero, J.C.; Harfouche, A.H.; Scarascia Mugnozza, G.; Moshelion, M.; Tuskan, G.A.; Keurentjes, J.J.B.; Altman, A. Accelerating Climate Resilient Plant Breeding by Applying Next-Generation Artificial Intelligence. Trends Biotechnol. 2019, 37, 1217–1235. [Google Scholar] [CrossRef]
  250. LeCun, Y.; Bengio, Y.; Hinton, G. Deep Learning. Nature 2015, 521, 436–444. [Google Scholar] [CrossRef]
  251. Liu, Y.; Wang, D.; He, F.; Wang, J.; Joshi, T.; Xu, D. Phenotype Prediction and Genome-Wide Association Study Using Deep Convolutional Neural Network of Soybean. Front. Genet. 2019, 10, 1091. [Google Scholar] [CrossRef]
  252. Voosen, P. How AI Detectives Are Cracking Open the Black Box of Deep Learning. Science 2017. [Google Scholar] [CrossRef]
  253. Hu, T.; Darabos, C.; Urbanowicz, R. Editorial: Machine Learning in Genome-Wide Association Studies. Front. Genet. 2020, 11, 593958. [Google Scholar] [CrossRef]
  254. Excoffier, L.; Lischer, H.E.L. Arlequin Suite Ver 3.5: A New Series of Programs to Perform Population Genetics Analyses under Linux and Windows. Mol. Ecol. Resour. 2010, 10, 564–567. [Google Scholar] [CrossRef]
  255. Li, H. A Statistical Framework for SNP Calling, Mutation Discovery, as- Sociation Mapping and Population Genetical Parameter Estimation from Sequencing Data. Bioinformatics 2011, 27, 2987–2993. [Google Scholar] [CrossRef] [Green Version]
  256. Maclean, C.A.; Hong, N.P.C.; Prendergast, J.G.D. Hapbin: An Efficient Program for Performing Haplotype-Based Scans for Positive Selection in Large Genomic Datasets. Mol. Biol. Evol. 2015, 32, 3027–3029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Goudet, J. HIERFSTAT, a Package for R to Compute and Test Hierarchical F-Statistics. Mol. Ecol. Notes 2005, 5, 184–186. [Google Scholar] [CrossRef] [Green Version]
  258. Manichaikul, A.; Mychaleckyj, J.C.; Rich, S.S.; Daly, K.; Sale, M.; Chen, W. Robust Relationship Inference in Genome-Wide Association Studies. Bioinformatics 2010, 26, 2867–2873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Chang, C.C.; Chow, C.C.; Tellier, L.C.A.M.; Vattikuti, S.; Purcell, S.M.; Lee, J.J. Second-Generation PLINK: Rising to the Challenge of Larger and Richer Datasets. Gigascience 2015, 4, s13742-015. [Google Scholar] [CrossRef] [PubMed]
  260. Pfeifer, B.; Wittelsbu, U.; Ramos-onsins, S.E.; Lercher, M.J. PopGenome: An Efficient Swiss Army Knife for Population Genomic Analyses in R. Mol. Biol. Evol. 2014, 31, 1929–1936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Kofler, R.; Orozco-terwengel, P.; De Maio, N.; Pandey, R.V.; Nolte, V.; Kosiol, C.; Schlo, C. PoPoolation: A Toolbox for Population Genetic Analysis of Next Generation Sequencing Data from Pooled Individuals. PLoS ONE 2011, 6, e15925. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Gautier, M.; Klassmann, A.; Vitalis, R. REHH 2.0: A Reimplementation of the R Package REHH to Detect Positive Selection from Haplotype Structure. Mol. Ecol. Resour. 2017, 17, 78–90. [Google Scholar] [CrossRef]
  263. Szpiech, Z.A.; Hernandez, R.D. Selscan: An Efficient Multithreaded Program to Perform EHH-Based Scans for Positive Selection. Mol. Biol. Evol. 2014, 31, 2824–2827. [Google Scholar] [CrossRef] [Green Version]
  264. Vilella, A.J.; Blanco-Garcia, A.; Hutter, S.; Rozas, J. VariScan: Analysis of Evolutionary Patterns from Large-Scale DNA Sequence Polymorphism Data. Bioinformatics 2005, 21, 2791–2793. [Google Scholar] [CrossRef] [Green Version]
  265. Danecek, P.; Auton, A.; Abecasis, G.; Albers, C.A.; Banks, E.; Depristo, M.A.; Handsaker, R.E.; Lunter, G.; Marth, G.T.; Sherry, S.T.; et al. The Variant Call Format and VCFtools. Bioinformatics 2011, 27, 2156–2158. [Google Scholar] [CrossRef] [PubMed]
  266. Kang, H.M.; Sul, J.H.; Service, S.K.; Zaitlen, N.A.; Kong, S.-Y.; Freimer, N.B.; Sabatti, C.; Eskin, E. Variance Component Model to Account for Sample Structure in Genome-Wide Association Studies. Nat. Genet. 2010, 42, 348–354. [Google Scholar] [CrossRef] [Green Version]
  267. Yang, J.; Lee, S.H.; Goddard, M.E.; Visscher, P.M. GCTA: A Tool for Genome-Wide Complex Trait Analysis. Am. J. Hum. Genet. 2011, 88, 76–82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Moser, G.; Lee, S.H.; Hayes, B.J.; Goddard, M.E.; Wray, N.R.; Visscher, P.M. Simultaneous Discovery, Estimation and Prediction Analysis of Complex Traits Using a Bayesian Mixture Model. PLoS ONE 2015, 11, e1004969. [Google Scholar] [CrossRef] [PubMed]
  269. do Prado Paim, T.; Hay, E.H.A.; Wilson, C.; Thomas, M.G.; Kuehn, L.A.; Paiva, S.R.; Mcmanus, C.; Blackburn, H. Genomic Breed Composition of Selection Signatures in Brangus Beef Cattle. Front. Genet. 2020, 11, 710. [Google Scholar] [CrossRef]
  270. Francis, J.; Little, D.A. Resistance of Droughtmaster Cattle to Tick Infestation and Babesiosis. Aust. Vet. J. 1964, 40, 247–253. [Google Scholar] [CrossRef]
  271. O’Neill, C.; Swain, D.; Kadarmideen, H. Evolutionary Process of Bos Taurus Cattle in Favourable versus Unfavourable Environments and Its Implications for Genetic Selection. Evol. Appl. 2010, 3, 422–433. [Google Scholar] [CrossRef]
Figure 1. Map of annual temperature changes in the period 1992–2020 compared to 1950–1978, created using ERA5 climate Reanalysis tools. The areas showing warming are in yellow-red and those showing cooling are in blue.
Figure 1. Map of annual temperature changes in the period 1992–2020 compared to 1950–1978, created using ERA5 climate Reanalysis tools. The areas showing warming are in yellow-red and those showing cooling are in blue.
Animals 11 02833 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Passamonti, M.M.; Somenzi, E.; Barbato, M.; Chillemi, G.; Colli, L.; Joost, S.; Milanesi, M.; Negrini, R.; Santini, M.; Vajana, E.; et al. The Quest for Genes Involved in Adaptation to Climate Change in Ruminant Livestock. Animals 2021, 11, 2833. https://doi.org/10.3390/ani11102833

AMA Style

Passamonti MM, Somenzi E, Barbato M, Chillemi G, Colli L, Joost S, Milanesi M, Negrini R, Santini M, Vajana E, et al. The Quest for Genes Involved in Adaptation to Climate Change in Ruminant Livestock. Animals. 2021; 11(10):2833. https://doi.org/10.3390/ani11102833

Chicago/Turabian Style

Passamonti, Matilde Maria, Elisa Somenzi, Mario Barbato, Giovanni Chillemi, Licia Colli, Stéphane Joost, Marco Milanesi, Riccardo Negrini, Monia Santini, Elia Vajana, and et al. 2021. "The Quest for Genes Involved in Adaptation to Climate Change in Ruminant Livestock" Animals 11, no. 10: 2833. https://doi.org/10.3390/ani11102833

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop