Go to Top Go to Bottom
Anim Biosci > Volume 35(11); 2022 > Article
Hariyono and Prihandini: Association of selected gene polymorphisms with thermotolerance traits in cattle – A review

Abstract

Thermal stress due to extreme changes in the thermal environment is a critical issue in cattle production. Many previous findings have shown a decrease in feed intake, milk yield, growth rate, and reproductive efficiency of cattle when subjected to thermal stress. Therefore, selecting thermo-tolerant animals is the primary goal of the efficiency of breeding programs to reduce those adverse impacts. The recent advances in molecular genetics have provided significant breeding advantages that allow the identification of molecular markers in both beef and dairy cattle breeding, including marker-assisted selection (MAS) as a tool in selecting superior thermo-tolerant animals. Single-nucleotide polymorphisms (SNPs), which can be detected by DNA sequencing, are desirable DNA markers for MAS due to their abundance in the genome’s coding and non-coding regions. Many SNPs in some genes (e.g., HSP70, HSP90, HSF1, EIF2AK4, HSBP1, HSPB8, HSPB7, MYO1A, and ATP1A1) in various breeds of cattle have been analyzed to play key roles in many cellular activities during thermal stress and protecting cells against stress, making them potential candidate genes for molecular markers of thermotolerance. This review highlights the associations of SNPs within these genes with thermotolerance traits (e.g., blood biochemistry and physiological responses) and suggests their potential use as MAS in thermotolerant cattle breeding.

INTRODUCTION

Recently, global climate change has had an adverse impact on the adaptability and survivability of farm animals to thermal assault. The elevation of ambient temperature, especially in summer, is a critical issue in the agricultural-related industry. One of the most important effects of thermal stress in the livestock industry is the decrease in cattle productivity and reproductivity. Numerous studies have demonstrated the significant decline in cattle performance when subjected to thermal stress; reduction in feed intake [1,2], milk yield [3,4], growth rate [5,6], and reproductive efficiency [7,8]. Thermal stress, such as heat stress, has also caused economic losses of around $897 million and $369 million annually for dairy and beef cattle industries, respectively [9], pressuring livestock producers to adopt effective preventative measures.
To mitigate the adverse impacts of thermal stress on cattle productivity, some strategies can be implemented, including genetic selection, modification of the environment, and nutritional intervention [10]. Besides, scientific techniques such as cooling systems and sprinkler treatments have been developed to reduce the adverse impact of heat stress on animals [11,12]. It has been reported that within-breed genetic variation exists for thermotolerance, allowing genetic selection for improved thermotolerance and may increasing animal resilience and welfare [13]. Furthermore, animal genetic selection for thermotolerance provides cumulative and permanent solutions at a relatively low cost [14]. To date, genetic selection in beef and dairy cattle is directed to productive traits, such as growth rate, meat production, milk yield, and milk quality. However, a continual selection for productive traits ignoring heat tolerance, results in decreasing heat tolerance [7]. There are considerable evidence that there is a genetic antagonism between productive traits and the specific ability to respond to heat stress [7,14,15]. For instance, selection for milk yield results in a poor response of the animals to heat stress [15]. Therefore, a combined selection for production and thermotolerance under highly modified environments, in both tropical and sub-tropical regions, should be considered. Furthermore, a set of phenotypes, including body temperature (BT) and respiration rate (RR), in combination with measures of indicators of heat tolerance in milk, can be used as an alternative in the genomic selection of heat-tolerant dairy cattle without damaging progress in milk production [14].
In conventional breeding methods, the selection of superior animals with economically important traits related to thermal stress is based on observed phenotypes and information on animal pedigree. However, the advancement of molecular genetics offers several approaches enabling the researchers to improve the accuracy of estimated breeding values and accelerate the achievements by selecting animals based on their genotypes rather than waiting to measure their physical attributes (phenotype) later in life. In molecular breeding methods, marker-assisted selection (MAS) is one of the approaches that allow the identification of specific DNA variations that are associated with particular economic traits like thermotolerance. Due to some advantages, single nucleotide polymorphisms (SNPs) are desirable DNA markers for MAS in livestock breeding. The SNPs are abundant throughout an organism’s genome (both coding and non-coding regions). The SNPs that cause an amino acid substitution in the gene coding regions are called non-synonymous; those that do not are synonymous. In addition, SNPs are more stably inherited than other DNA markers, making them more suited as long-term selection markers [16].
In cattle, several SNPs within some genes (e.g., HSP70, HSP90, HSF1, EIF2AK4, HSBP1, HSPB8, HSPB7, MYO1A, and ATP1A1) have been reported to be associated with thermotolerance traits [1725]. All the mentioned genes have critical roles in many cellular activities during thermal stress, protecting cells against stress. Furthermore, stress-induced gene expression and activation of heat shock proteins (HSPs) are key indicators of the animal’s response to thermal stress in cellular and molecular activities [26]. Some potential indicators of thermal stress, such as blood biochemistry and physiological responses (BT, RR, and rectal temperature [RT]) were proposed to elucidate the impact of the mentioned genes on thermotolerance in cattle [17,18,27]. BT and RR can be measured using automated phenotyping technologies [28]. Therefore, it is very important to have deep knowledge about the genetic basis of the animal’s response to thermal stress to identify molecular markers for thermotolerance in livestock breeding.
This study discussed some potential candidate genes and their associations with thermotolerance traits in cattle. The genes discussed in this review included HSP70, HSP90, HSF1, EIF2AK4, HSBP1, HSPB8, HSPB7, MYO1A, and ATP1A1 (Table 1).

THERMOTOLERANCE TRAITS IN CATTLE

It is well-known that temperature is one of the significant environmental factors affecting the livestock populations’ health and productivity. Rising global temperatures put heat stress at the center of ever-growing concern regarding livestock populations across tropical and temperate zones [29]. Genetic selection for thermotolerance is likely to be one of the potential strategies to overcome the effect of rising temperatures on cattle productivity. Carabaño et al [30] proposed several measurements as indicators of thermotolerance in animals, including BT, heart rate (HR), RR, and sweating rate (SR). Several biomarkers like blood biochemical parameters have also been proposed as selection criteria to identify thermo-tolerant animals [17]. It has been reported that RT is the best predictor of core BT. The heritability of RT during heat stress is moderate making it possible to use this trait as selection criteria to improve thermotolerance [31]. However, the inheritance of physiological traits, such as RT and HR is still little known, owing the fact that they are extremely labor-intensive to collect, especially in a large-scale selection program. Koltes et al [28] proposed some automated temperature detection devices that can be used to provide real-time temperature data and monitor potential thermal stress in animals. Those devices allow research that can inform producers on how best to manage or select thermo-tolerant animals.
Identification of thermo-tolerant animals is challenging because the responses to heat stress are complex and variable. Besides, there is antagonism between thermotolerance and productive traits, as reported in some dairy cattle populations [7,14,15,32]. For instance, selection for higher milk yield results in a reduced animal’s ability to cope with heat stress [15]. Negative genetic correlations from −0.85 to −0.24 between thermotolerance and productive traits (e.g., milk yield) have been observed in dairy cattle [32]. Therefore, it is important to monitor thermotolerance when selection for productive traits is implemented in production systems influenced by heat stress. Furthermore, there are substantial genetic variations in an individual animal to cope with greater heat loads, with a moderate degree of genotype x environment interaction. This suggests that animals that produce the most in comfort may not be the best animals under heat stress [30]. Consequently, the genetic variation in tolerance to heat stress should be considered in the selection of dairy cattle raised under modified environments in both tropical and sub-tropical regions.

CANDIDATE GENES FOR THERMOTOLERANCE

Heat shock protein 70 (HSP70)

Heat shock proteins (HSPs) are a group of proteins that protect cells from harmful oxidative stress. The HSPs family consists of HSP110, HSP100, HSP90, HSP70, HSP60, HSP40, HSP10, and small HSP [33]. Among the HSPs, the proteins of molecular mass 70 kDa, termed HSP70, is the most essential molecular chaperone of primary importance to all mammalian cells, which acts by binding to other cellular proteins and facilitating intracellular transport [34]. HSP70 is produced by the HSP70 gene, characterized by a single exon. The open reading frame of this gene is approximately 1926, and its protein consists of 641 amino acids, of which 92 and 82 are highly basic and acidic amino acids, respectively [35]. Heat shock genes are activated by stressful stimuli, further forming HSPs. The HSP70 expression in heat-stressed dairy animals was found to be the most responsive compared to other HSPs chaperones. Therefore, the HSP70 expression profile could be beneficial as a biomarker to determine the effect of thermal stress in animals [36].
The HSP70 gene family in bovine consists of four genes, including HSP70-1, HSP70-2, HSP70-3, and HSP70-4 [37]. In cattle, both the HSP70-1 and HSP70-2 genes are located on chromosome 23 band 22 (the BoLA region). In contrast, the HSP70-3 gene is located on chromosome 10 band 34, and the HSP70-4 is localized at chromosome 3 band 13 [38]. Gallagher et al [38] screened bovine genomic DNA for hybridization with the HSP70 cDNA in humans. They found that bovine HSP70-1,2 were homologous with human HSPA1 and HSPA1L on chromosome 6p21.3, while bovine HSP70-3 was homologous with a human HSP70 gene on chromosome 14q22-q24 and bovine HSP70-4 was the homolog of the human HSPA-6,-7 genes on chromosome 1.
Figure 1 shows the molecular mechanism for the expression of HSP70 within a cell. A cell undergoes molecular changes when exposed to thermal stress. In this regard, most of the HSP family genes, including HSP70, have crucial cytoprotective effects and are involved in many regulatory pathways related to cell stress response [33]. Besides the behavioral and physiological responses, thermal stress can induce the HSP70 expression [39,40]. However, genetic variation among individual animals may cause varying thermal stress coping capacities. Therefore, understanding the physiological systems that are controlled by genes involved in thermoregulation is important to prioritize variants in genomic selection strategies in addition to understanding the modulation of genotype × environment interaction [41].
Xu et al [42] identified 193 differentially expressed genes in peripheral blood samples of Sanhe cattle exposed to severe cold stress (−32°C for 3 h). Out of the 193 candidate genes, the expression of HSP70 significantly increased after cold exposure (p<0.05). An earlier study conducted by Abbas et al [17] on lactating Holstein cows under heat and cold stress observed a significant association between SNPs in the 5′ flanking region of the HSP70 gene and blood biochemical parameters. SNPs A-12G and C181T in the HSP70 gene significantly affected (p<0.05) lactate concentration, while A72G was associated (p<0.05) with lipid peroxide concentration under heat stress conditions, and the same association exists for SNPs A-12G and SNP C131G with dopamine and superoxide dismutase concentrations [17].
Polymorphisms in the HSP70 gene have also been associated with blood biochemistry in Inner-Mongolia Sanhe cattle, a dual-purpose (milk and meat) cattle breed in the Inner Mongolia Autonomous Region of China [43]. Among 20 SNPs identified in the HSP70 gene in Sanhe cattle, SNP-42 located in the promoter region and SNPs −105+, −181+, and −205+ localized in the 5′-untranslated region (UTR) have been associated with blood triiodothyronine (T3), while SNP −105+ has been associated with blood thyroxine (T4). The SNPs −42 and −205+ have been observed to be the causative mutations involved in regulating HSP70 promoter activity [43]. Amino acid changes from aspartate to tyrosine and from aspartate to histidine due to G>T and G>C substitutions, respectively, at 149th of the amplicon (295 bp) in the HSP70 gene in Tharparkar cattle have been identified to produce two alleles (alleles A and B with nucleotides T and C, respectively), in which genotype AA was superior in heat tolerance [44]. Also, the identified SNP was novel and has been submitted to National Center for Biotechnology Information (NCBI) with GenBank Accession Number JX966362 for allele A and JX966363 for allele B. Onasanya et al [45] in their earlier study identified twelve SNPs of the HSP70 gene in four zebu breeds of Nigeria; White Fulani, Sokoto Gudali, Red Bororo, and Ambala. Of the twelve variants, four (C151T, C146T, G90A, and C219A) were unique as they were detected in all the analyzed breeds. The SNP C146T was synonymous, while C151T, G90A, and C219A were non-synonymous mutations, causing changes in the coded proteins from serine to leucine, from alanine to arginine, and from cysteine to threonine, respectively. Regarding the four SNPs, heterozygous animals had lower a heat tolerance coefficient (HTC), suggesting their potential to withstand heat stress than homozygous counterparts [45].
Prihandini et al [46] identified fourteen SNPs in the 5’-UTR of the HSP70 gene in several cattle breeds in Indonesia (Galekan, Bali, Jabres, Belgian Blue×Peranakan Ongole [PO] cross, Rambon, Madura, and PO). The SNPs included C1036T, G1045A, A1058G, C1069T, G1076A, A1096G, G1117A, A1125C, G1128T, T1134C, G1164T, T1204C, C1255T, and C1262T. Among these SNPs, SNP G1045A, T1134C, and T1204C have been associated with the serum concentration of T3 and IGF-I and body condition [47,48], while the SNP A1125C was associated with calving percentages in crossbred Brahman cows [48]. The SNP G1076A identified by Prihandini et al [46] in Indonesian cattle breeds was also observed in South African Nguni crossbred cattle [49]. Six SNPs in the HSP70 gene, including G1045A, G1117A, A1125C, G1128T, T1134C, and T1204C, have been identified in some cattle populations from different countries [46,5052]. Of the six SNPs, SNP G1128T has been associated with cell viability and gene expression of the HSP70.1 gene in Holstein cattle, in which heterozygous animals showed higher cell viability (p<0.05) and gene expression (p<0.001) compared to homozygous GG [50]. The SNP G1128T was also associated with calving percentages in crossbred Brahman cows [48]. Badri et al [53] found SNP G462T in the coding region of the HSP70 gene in Chinese Holstein cattle with an amino acid substitution at position 154 from glutamine to histidine, which may influence heat tolerance.

Heat shock protein 90 (HSP90)

HSP90 is a member of the HSPs family, which has an important biological role in protein translocation and regulation of steroid hormone receptors [36]. The HSP90 is a versatile chaperone protein that maintains cellular integrity and homeostasis during heat stress [39]. This protein is considered a highly conserved and essential stress protein expressed in all eukaryotic cells, comprising 1% to 2% of cellular proteins under non-stress conditions [54]. The HSP90 is the regulator for nearly 100 proteins termed client proteins, involved in signal transduction [54]. Among the HSPs, HSP90 and HSP70 are the two important components of the cellular machinery involved in protein homeostasis and participate in nearly every cellular process. In some chaperone activities the two proteins work together in cellular remodeling functions [55] and act as apoptosis modulators, by interfering with the formation of the apoptosome, thereby reducing cell death during heat shock [56].
When the mammalian cells are exposed to external stress, a way to respond to such stress is by releasing some conserved proteins, such as the HSPs family. Among the HSPs, the HSP90 expression has been associated with the freezing resistance of bull sperm. The higher expression of HSP90 leads to the higher motility and freezing resistance of sperm after freezing-thawing [57]. HSP90 expression in bull spermatozoa gradually declined following freezing-thawing process, thereby affecting sperm quality [58]. There are two major cytoplasmic isoforms of HSP90, including the inducible form (HSP90α or HSP90AA1) and the constitutive form (HSP90β or HSP90AB1), which are constituted by the duplication of genes [27,59]. HSP90AA1 and HSP90AB1 genes have been mapped to bovine chromosome 21 (with 733 amino acids) and 23 (with 724 amino acids), respectively. The HSP90 isoforms contribute to cellular functions, including signal transduction, protein folding, protein degradation, cell survival, and morphological evolution [60].
Certain polymorphisms in the HSP90 gene in various breeds of cattle throughout the world have been identified. Onasanya et al [18] identified four SNPs in the exon 3 of the HSP90 gene in four zebu breeds in Nigeria, i.e., Ambala, Red Bororo, Sokoto Gudali, and White Fulani. The identified SNPs were T116G, G220C, G346A, and G390A. Heterozygous animals showed significantly (p<0.0001) lower BT, RT, RR, and HTC than homozygous animals [18]. The HSP90AB1 gene polymorphisms have also been detected by Charoensook et al [61] in indigenous Thai breed White Lamphun and Mountainas, as well as a crossbred Holstein Friesian. In their studied populations, a total of nine SNPs were observed, one in the 3’-UTR (g.5435T>C), three in exons 10 (g.4374T>G) and 11 (g. 5007T>C and 5082C>T), and five in introns 8 (g. 4029G>C and 4061G>A), 9 (g.4338T>C), 10 (g.4730A>C), and 11 (g.5248C>T). The SNP g. 5082C>T in exon 11 results in an alanine-to-valine amino acid change. Furthermore, the T allele at SNP T4338C in intron 9 significantly improved heat tolerance (p<0.05) [61]. Interestingly, the SNP g.4338T>C has also been detected in Indian breeds of dairy cattle, Sahiwal and Frieswal, in which TT genotype had significantly (p< 0.01) higher HTC than CT and CC genotypes [62].
Genetic analysis of the HSP90AA1 gene in Karan Fries (5/8 Holstein Friesian × 3/8 Tharparkar) showed a point mutation, g.1209A>G in exon 3, which produced three genotypes, i.e., AA, AG, and GG. Association analysis for the SNP g.1209A>G showed that homozygous GG had significantly (p<0.01) lower RR, RT, and HTC than homozygous AA and heterozygous AG [27]. A total of five SNPs of the HSP90AA1 gene were detected in Chinese Holstein cows; one in the promoter (g.−87G>C), three in the coding (g.605A> G, 1662T>G, 2819G>A), and one in 3′-UTR (g.4172A>G) regions [63]. For the g.−87G>C locus, animals having CC genotype had significantly (p<0.01) higher HSP90AA1 mRNA expression than those having GG genotype in stress conditions [63]. Thus, the SNPs detected in the HSP90 gene could be potential molecular markers for thermotolerance in cattle breeding.

Heat shock factor 1 (HSF1)

Heat shock factors (HSFs) are the main transcription factors that maintain protein homeostasis (proteostasis) and counteract disturbances by regulating heat shock gene transcription [64]. The HSFs are activated in response to stress factors, such as rising temperature, while inactive in non-stressed cells [65]. The HSFs transiently activate the HSPs transcription by binding to their promoters [64]. The HSPs were first observed in Drosophila melanogaster in the early 1960s, as revealed by a “puffing” of genes in the chromosome of recovering cells [66]. The “puffing” has been shown to activate genes encoding the HSPs, which function as molecular chaperones [67]. The HSF family consists of four members in mammals, including HSF1, HSF2, HSF3, and HSF4. Among them, HSF1 is the primary regulator of the heat shock gene transcription in response to stress conditions [68]. The HSF1 gene was mapped to bovine chromosome 14, and appears to be involved in cardiac health, cell functions, stress responses, and DNA damage repair in mammals [19,69]. Animals with an abundance of the HSF1 gene expression have a great capacity to cope with heat stress compared to those with lower expression [70]. Consequently, the HSF1 expression can be used as indicator of thermotolerance in cattle.
Although the importance of the HSF1 gene as a regulator of the heat shock responses, a few studies have been conducted on identifying the HSF1 polymorphisms in cattle when compared to the HSPs family [19,71,72]. Sharma et al [73] characterized the 3’-UTR of the HSF1 gene in Indian cattle breeds, consisting of three SNPs (G1762T, C1811T, and C1983T), and binding sites for several miRNAs. They also noted the abundance expression of most of the miRNAs after heat stress in peripheral blood mononuclear cells of the studied animals, which may affect the function of the HSF1 gene in response to heat stress. A total of twelve SNPs of the HSF1 gene were identified in Angus cattle raised in subtropical conditions [72]. Of the twelve, eight were in coding regions with no alteration in the amino acid sequence of the protein, and others were mapped to intron regions with uncharacterized function. Rong et al [19] identified SNP NC_037341.1 g.616087A>G (rs135258919) in the HSF1 gene in 35 Chinese cattle breeds, which produced three genotypes, AA, AG, and GG. The SNP was associated with annual temperature (T), relative humidity (RH), and temperature-humidity index (THI) (p<0.01). The mutant allele (G) of the SNP was highly detected in animals raised in regions with more prominent heat stress, suggesting that animals carrying allele G (GG and AG genotypes) may be more tolerant to heat stress than those with allele A (AA genotype) [19]. The genetic polymorphisms of the HSF1 gene were also studied in Chinese Holstein cattle, in which two SNPs (T909C in intron 3 and G4693T in 3′UTR) significantly affected thermotolerance [71]. Based on these SNPs, nine haplotype combinations were constructed. Animals carrying H2H4 haplotype combination (TCTT) showed higher HTC (p<0.05) and lower potassium content in erythrocytes (p<0.01), decreased rate of milk production (p<0.05), and RT (p<0.05) than those carrying H1H3 haplotype combination (TCGG) [71]. Li et al [71] further noted that the 4693-T mutation caused the disruption of microRNA target binding, resulting in the relief of the transcriptional repression that led to the increased HSF gene expression [71]. Furthermore, the SNP G4693T was noted to affect the HSF1 gene expression by influencing the binding of HSFI to bta-miR-484 [74]. Li et al [74] further observed another SNP G1451T, in intron 3 of the HSF1 gene. Still, this SNP was noted to have no effect on thermotolerance.

Eukaryotic translation initiation factor 2-alpha kinase 4 (EIF2AK4)

The alpha subunit of eukaryotic initiation factor -2 (eIF2α) is a molecule involved in regulating protein synthesis initiation in eukaryotes [75]. The eIF2α is phosphorylated by protein kinases, activated in response to various stress conditions; oxidative stress, heme deficiency, osmotic shock, and heat shock [76]. Eukaryotes have four types of eIF2α kinases, including PKR, HRI, PERK, and EIF2AK4 [20]. Among these, the EIF2AK4 gene represses the translation of most mRNAs in response to stress-induced signals [77]. The EIF2AK4 gene is also known as the general control non-derepressible 2 (GCN2) gene. This gene is activated in response to viral infection and oxidative stress [78,79]. The activation of the EIF2AK4 gene also occurs by binding uncharged transfer RNAs (tRNAs) to the HisRS domain of the protein [80].
Heat exposure in mammals can cause oxidative stress and DNA damage resulting in apoptosis [81]. Edea et al [69] detected candidate genes associated with thermo-tolerance and DNA damage repair, including HSPA4, HSF1, CMPK1, and EIF2AK4. The EIF2AK4 gene plays a central role in regulating gene expression in response to the deprivation of amino acids and glucose [82]. Moreover, the EIF2AK4 gene was noted to activate and mediate cellular response to DNA damaging agents (e.g., UV), viral infections, and nutritional deprivation [83]. A recent study by Wang et al [20] characterized the EIF2AK4 gene in 35 Chinese cattle breeds and two exotic breeds (Angus [Bos taurus] and Burma [Bos indicus]). The investigated Chinese cattle breeds were classified into southern cattle, which genetically belonged to indicine cattle, and northern cattle groups, which were genetically classified as taurine cattle. They noted a novel SNP, NC_037337.1 g.35615224 T>G, located in exon 6 of the EIF2AK4 gene, responsible for a p.Ile205Ser substitution. The frequency of the wild-type allele T (NC_037337.1 g.35615224 T>G) gradually increased from the southern cattle groups to the northern cattle groups, while the mutant allele G showed an opposite pattern. The SNP was also associated with thermotolerance in Chinese cattle, as revealed by the GG genotype, which was generally found in regions with higher AT, RH, and THI [20]. However, the mechanisms by which the EIF2AK4 gene expression affects cattle thermotolerance must be further investigated. As noted by Jiang and Wek [84], the EIF2AK4 gene was also associated with resistance to apoptosis due to UV irradiation. Loss of EIF2AK4 gene significantly increases apoptosis during UV exposure. Furthermore, the EIF2AK4 gene plays a pivotal role in the integrated stress response (ISR) by regulating sensing starvation [85].

Heat shock factor binding protein 1 (HSBP1)

Heat shock factor binding protein 1 (HSBP1) is a novel and conserved protein containing 76 amino acids and two extended arrays of hydrophobic repeats [86]. The HSBP1 interacts with the oligomerization domain of the HSF1 to suppress HSF1's transcriptional activity in response to heat shock [86, 87]. Besides, the HSBP1 also interacts with HSP70 at a later point in the heat shock response relative to the appearance of HSBP1/HSF1 complexes, inhibiting HSF’s capacity to bind DNA and conversion of the trimer to monomer state [86]. At 1.8 A resolution, the HSBP1 contained crystal structure and amino acid residues 6–53, forming a continuous 11-turn long helix [88]. The HSBP1 gene acts as a negative regulator of heat shock responses, thereby affecting the survival of animals when subjected to thermal stress [86]. In mice, although the HSBP1 reduction in embryoid bodies (Ebs) leads to increased HSF1 activity, it causes defects in the organization of the germ layers and a reduction in the expression of definitive endodermal markers [87]. In cattle, genetic polymorphisms of the HSBP1 gene were noted on 930 Chinese Holstein cattle [21]. Consequently, three SNPs (e.g., g.324G> C, g.589C>T, and g.651C>G) formed seven haplotypes and fourteen haplotype combinations. Among the fourteen, the H2H2 haplotype combination had a lower decrease rate of milk yield than the H2H3 haplotype combination (p<0.05). Also, lower potassium content in erythrocytes (PCE) was noted in H2H2 haplotype combination as compared to H2H5 (p<0.05), H4H4 (p<0.05), and H4H5 (p<0.01) haplotype combinations. Moreover, it was noted that the SNP g.651C>G affected thermotolerance in the analyzed cattle; cows carrying GG genotype had lower PCE than CG genotype (p<0.01) [21].

Heat shock protein family B (small) member 8 (HSPB8)

Heat shock protein family B (small) member 8 (HSPB8), also known as HSP22, H11 kinase, or E2IG, is a member of the small HSP family, which is expressed in response to heat shock [89,90]. The HSPB8 is present in the plasma membrane and interacts with the lipid membrane to result in the burial of the tryptophan residues and observable conformational change, thereby interfering with cellular activities such as signal transduction and apoptosis [91]. HSPB8 tends to form small-molecular-mass oligomers and interacts with biological membranes and many different proteins, such as glycolytic enzymes and different protein kinases [92]. Moreover, this protein exists as a monomer in vitro and involves in chaperone-like activity as well [89]. The protein also acts as a positive regulator in the PGF2α-induced synthesis of interleukin-6 (IL-6) and vascular endothelial growth factor A (VEGF) in osteoblasts [93]. The HSPB8 gene has been mapped to bovine chromosome 17 (BTA17q24-25) [22]. In mammals, mutations in the gene encoding HSPB8 can lead to the development of various diseases such as myopathies and neuropathies [90]. The HSPB8 expression has also been associated with mitochondrial dysfunction that leads to subarachnoid hemorrhage (SAH)-induced early brain injury (EBI) [94]. Verma et al [22] analyzed 108 Sahiwal indigenous cattle in India to reveal SNPs within the HSPB8 gene and their association with thermotolerance. Two SNPs (e.g., g.507G>A in exon 1 and g.881T>C in intron 1) were noted, but only SNP g.507G>A significantly affected heat tolerance in the studied cattle. For g.507G>A locus, animals with GA genotype had significantly lower (p<0.01) RR, RT, and HTC than those with GG genotype. Therefore, the GA genotype of SNP g.507G>A of the HSPB8 gene may be advantageous in selecting heat-tolerant animals [22].

Heat shock protein family B (small) member 7 (HSPB7)

The HSPB family is one of the most diverse families within the group of HSP families [95]. The HSPB7 is a member of the HSPB, which is preferentially expressed in cardiovascular, skeletal muscle, and adipose tissue [96]. In obese rats, the HSPB7 mRNA expression increases in skeletal muscle and brown and white adipose tissues [97]. The HSPB7 gene appears in oligomeric and dimeric forms (approximately 17 kDa and 40 kDa, respectively) composed of 18.6-kDa monomers [98]. Unlike HSPB1 and HSPB5, that chaperoned heat unfolded substrates and kept them folding competent, HSPB7 did not support refolding [95]. In cattle, the overexpression of HSPB7 improved the H2O2-induced oxidative stress in adipocytes via increasing the abundance of NFE2 like bZIP transciption factor 2 (NFE212) and its downstream target genes heme oxygenase-1 (HMOX1) and NADH quinone oxidoreductase 1 (NQO1). Knockdown of HSPB7 markedly inhibited the expression of NFE2L2, HMOX1, and NQO1 and further exacerbated H2O2-induced oxidative stress [99]. In humans, HSPB7 knockdown promoted osteogenic differentiation of human adipose-derived stem cells (hASCs) via activation of the ERK1/2 signaling pathway [100]. Zeng et al [23] explored 774 individuals representing 32 Chinese indigenous cattle breeds to detect the polymorphism in the HSPB7 gene. Consequently, one SNP NC_037329.1: g.136054902 C>G was found, of which the allele C was dominant in northern cattle groups. In contrast, allele G was dominant in southern indicine cattle groups. The SNP was associated with AT, RH, and THI (p<0.01). Animals carrying the allele G had higher T, RH, and THI. Thus, SNP NC_037329.1: g.136054902 C>G might benefit heat-tolerant cattle breeding [23].

Myosin-1a (MYO1A)

Myosin-1a (MYO1A) gene is a candidate gene associated with skin pigmentation in cattle [69]. As reported previously, skin pigmentation is highly related to the BT regulation [101,102]. In the enterocyte microvillus, the MYO1A was considered the most abundant actin-based motor protein interacting with the apical membrane via a highly basic C-terminal tail domain [103]. The MYO1A belongs to one of eight monomeric, membrane-binding, and actin-based motor protein class I myosins, which are highly expressed in vertebrates [104]. It is also a mechanoenzyme previously thought to be located exclusively in the intestinal epithelium [105]. Furthermore, MYO1A plays an important role in the apical membrane movement and structural stability [106] by powering the sliding of the apical membrane along with microvillar actin bundles [107] and regulating membrane-cytoskeleton adhesion, which enables the apical membrane to resist deformation [108].
In cattle, an association of the MYO1A gene polymorphisms with thermotolerance was first described by Jia et al [24] on 1,072 animals from 34 Chinese indigenous cattle breeds, and Angus and Indian zebu. Their study found four SNPs within the MYO1A gene, including rs208210464, rs110123931, rs209999142, and rs135771836. A missense mutation of rs209999142 resulted in an amino acid change from phenylalanine to serine. All the identified SNPs were significantly associated with environmental parameters, including T, RH, THI, and SR. Besides, Hap 1/1 constructed based on the four SNPs was advantageous in selecting heat-tolerant animals [24]. Cao et al [109] reported a novel SNP (rs209559414 or NC_037332.1 g.56390345 A>G) within the MYO1A gene in Chinese indigenous cattle. This SNP was capable of resulting in an amino acid substitution from isoleucine into valine. The frequency of wild allele A decreased gradually from northern cattle (a temperate monsoon climate) to southern cattle (a subtropical monsoon climate), while that of mutant type allele G showed the opposite pattern. A significant association between genotypes for the SNP rs209559414 and climatic conditions (AT, RH, THI, and average annual sunshine hours [100-cloudiness]) was noted [109].

ATPase Na+/K+ transporting subunit alpha 1 (ATP1A1)

ATP1A1 gene is a well-known gene involved in regulating BT during heat stress in cattle [25,110112]. The ATP1A1 gene has been mapped to bovine chromosome 3, consisting of 22 introns and 23 exons [113]. This gene encoded the alpha-1 chain of Na+/K+-ATPase, an integral membrane protein of a heterodimeric enzyme, which mediates cutaneous vasodilation during heat stress by interacting with nitric oxide synthase (NOS) [114,115]. This protein also has an important role in membrane permeability by coupling the transport of three Na+ ions outward and two K+ ions inward [116]. The Na+/K+-ATPase consists of four isoforms of an α subunit (α1–α4), three isoforms of a β subunit (β1–β3), and FXYD proteins [117]. The four α ATPases include ATP1A1, ATP1A2, ATP1A3, and ATP1A4, encoding for α1, α2, α3, and α4 protein subunits, respectively [118]. Among the α ATPases, the ATP1A1 was highly expressed in erythrocytes and peripheral nerves [115]. It has been reported that the ATP1A1 mRNA expression of dairy cows under heat stress was higher than those under optimal temperature [110].
Numerous research has been undertaken on various breeds of cattle to determine the association between ATP1A1 gene polymorphisms and thermotolerance. Liu et al [25] observed two SNPs, −14103G>A in exon 14 and −14242C>T in intron 14 of the ATP1A1 gene in Holstein cows, which were significantly associated with HTC and RR (p<0.01). Animals with AC genotype were the most tolerant coping with heat stress. A year later, Liu et al [110] discovered a novel synonymous mutation (C2789A) in exon 17 of the ATP1A1 mRNA in Holstein dairy cows, significantly associated with heat resistance. Due to this SNP, cows carrying the CC genotype showed significantly higher heat tolerance than those carrying the CA genotype (p<0.05) [110]. This SNP was also identified in Vrindavani and Tharparkar cattle of India using polymerase chain reaction single-strand conformation polymorphism and DNA sequencing and associated with heat resistance. Association analysis for this SNP showed that animals with genotype CC showed higher HTC and lower RT than CA and AA genotypes in both breeds [112]. Furthermore, the TT genotype at T27008243C locus in Sahiwal and the AA genotype at 27008223 locus in Karan Fries cows were the most favorable genotypes for heat tolerance [111]. In Cholistani cattle, genetic variant (BB) of the ATP1A1 gene was found to have significant (p<0.05) effect on vaginal temperature [119]. The polymorphism of the ATP1A1 gene has also been associated with productive traits like feed intake in European beef cattle breeds [120].

CONCLUSION AND FUTURE PERSPECTIVES

There is a significant association of SNPs in HSP70, HSP90, HSF1, EIF2AK4, HSBP1, HSPB8, HSPB7, MYO1A, and ATP1A1 genes with thermotolerance in various breeds of cattle (both dairy and beef cattle), implying their potential uses as molecular markers in breeding schemes. These molecular markers may benefit both management and breeding decisions to select thermal-tolerant cattle. Since thermal stress is a critical issue in cattle production, mainly due to its deleterious impact on feed intake, milk yield, growth rate, and reproductive efficiency, knowledge of these candidate genes regulating thermal stress can be valuable in providing MAS in cattle breeding. Identification of SNPs in stress-related genes has been proved to generate beneficial molecular data in farm animals. It has been proposed that genomic selection is a promising approach to accelerate genetic gain for thermotolerance because young bulls and heifers can be selected based on their genomic estimated breeding value (GEBV) [32,121]. The higher the young bulls genotyped, the smaller the marginal cost of the additional GEBV is. However, the main challenge in developing a GEBV is the size of the reference population [122]. The accuracy of genomic prediction increases in line with the increased size of the reference population. In dairy cattle, genomic predictions for heat thermotolerance have been conducted using reference populations from Holstein and Jersey cattle genotyped for 632,003 SNPs with an accuracy of 0.39 to 0.57 in Holsteins, and 0.44 to 0.61 in Jerseys [32]. This suggests that genomic selection is an excellent alternative to improve thermotolerance in cattle. Furthermore, emerging technologies in molecular genetic techniques and genome editing, such as clustered regularly interspaced short palindromic repeat (CRISPR)/CRISPR-associated endonuclease Cas9 (Cas9), may pave the way for novel approaches, allowing introduction of site-specific gene modifications. Finally, incorporating molecular information into breeding programs will benefit industry, scientists, and breeders to develop thermo-tolerant cattle and improve the accuracy of traditional selection methods.

Notes

CONFLICT OF INTEREST

We certify that there is no conflict of interest with any financial organization regarding the material discussed in the manuscript.

FUNDING

The authors received no specific funding for this work.

Figure 1
Molecular mechanism for the expression of HSP70 within a cell, adapted from Kregel [40]. The HSP70 expression is modulated by some physiological signals like thermal stress, energy depletion, acidosis, and viral infection. These physiological stimuli activate the heat shock factors (HSFs), causing them to seperate from heat shock proteins (HSPs). The HSFs are phosphorylated by protein kinases and form trimers in the cytosol. These HSF trimer complexes enter the nucleus and bind to heat shock elements (HSE) in the promotor region of the HSP70 gene. HSP70 mRNA is then transcribed and leaves the nucleus for the cytosol, where new HSP70 is synthesized. Proposed mechanisms of cellular protection for HSPs include their functioning as molecular chaperones to assist in the assembly and translocation of newly synthesized proteins within the cell and the repair and refolding of damaged (e.g., stress-denatured) proteins.
ab-22-0055f1.jpg
Table 1
Effects of SNPs in some genes on thermotolerance in cattle
Gene SNP GenkBank accession number Location Amino acid substitution Breeds analyzed Associated trait Reference
HSP70 A-12G AY149618.1 5′ Flanking Region None Holstein cattle LA, DA,and SOD [17]
C181T AY149618.1 5′ Flanking Region None Holstein cattle LA [17]
A72G AY149618.1 5′ Flanking Region None Holstein cattle LPO [17]
C131G AY149618.1 5′ Flanking Region None Holstein cattle DA and SOD [17]
−42 (C/T) AY149618.1 Promoter None Inner-Mongolia Sanhe cattle Blood triiodothyronine (T3) [43]
−105+(G/T) AY149618.1 5′-UTR None Inner-Mongolia Sanhe cattle Blood triiodothyronine (T3) and thyroxine (T4) [43]
−181+(C/T) AY149618.1 5′-UTR None Inner-Mongolia Sanhe cattle Blood triiodothyronine (T3) [43]
−205+(C/T) AY149618.1 5′-UTR None Inner-Mongolia Sanhe cattle Blood triiodothyronine (T3) [43]
G149T NM_174550.1 Coding region Aspartate to Tyrosine Tharparkar cattle RT, ART, HTC, and RR [44]
G149C NM_174550.1 Coding region Aspartate to Histidine Tharparkar cattle RT, ART, HTC, and RR [44]
C146T NM_174550.1 Exon 1 None Nigerian zebu: White Fulani, Sokoto Gudali, Red Bororo, and Ambala HTC [45]
C151T NM_174550.1 Exon 1 Serine to Leucine Nigerian zebu: White Fulani, Sokoto Gudali, Red Bororo, and Ambala HTC [45]
G90A NM_174550.1 Exon 1 Alanine to Arginine Nigerian zebu: White Fulani, Sokoto Gudali, Red Bororo, and Ambala HTC [45]
C219A NM_174550.1 Exon 1 Cysteine to Threonine Nigerian zebu: White Fulani, Sokoto Gudali, Red Bororo, and Ambala HTC [45]
G1128T U09861 Promoter region None Holstein cattle Cell viability, HSP70 expression [50]
HSP90 T116G AC_000178.1 Exon 3 Threonine to Histidine Red Bororo cattle BT, RT, RR, and HTC [18]
G220C AC_000178.1 Exon 3 Arginine to Serine Sokoto Gudali BT, RT, RR, and HTC [18]
G346A AC_000178.1 Exon 3 Serine to Leucine Ambala BT, RT, RR, and HTC [18]
G390A AC_000178.1 Exon 3 Aspartate to Tyrosine White Fulani BT, RT, RR, and HTC [18]
HSP90AB1 T4338C NW001494158 Intron 9 None White Lamphun and Mountainas HTC [61]
T4338C NW001494158 Intron 9 None Sahiwal and Frieswal HTC [62]
C5248T NW001494158 Intron 11 None White Lamphun and Mountainas HTC [61]
C1787061T Not mentioned Intron 10 None Sahiwal RT [123]
HSP90AA1 g.1209A>G AC_000178.1 Exon 3 None Karan Fries RR, RT, and HTC [27]
g.−87G>C AC_000178.1 Promoter region None Chinese Holstein HSP90AA1 mRNA expression [63]
HSF1 g.616087A>G NC_037341.1 Coding region Valine into Alanine Chinese cattle T, RH, THI [19]
T909C NC_007312.2 Intron 3 None Chinese Holstein PCE [71]
G4693T NC_007312.2 3’-UTR None Chinese Holstein HTC, HSF1 expression [71]
G1451T NC_007311.3 Intron 3 None Chinese Holstein None [74]
EIF2AK4 g.35615224 T>G, NC_037337.1 Exon 6 Isoleucine to Serine Chinese cattle, Burma, Angus T, RH, and THI [20]
HSBP1 g.324G>C NC_007316.3 Intron 1 None Chinese Holstein None [21]
g.589C>T NC_007316.3 Intron 1 None Chinese Holstein None [21]
g.651C>G NC_007316.3 Exon 2 None Chinese Holstein PCE [21]
HSPB8 g.507G>A AC_000174.1 Exon 1 None Sahiwal RR, RT, and HTC [22]
g.881T>C AC_000174.1 Intron 1 None Sahiwal None [22]
HSPB7 g.136054902 C>G NC_037329.1 Coding region Alanine to Glycine Chinese cattle, Angus, Indian zebu T, RH, and THI [23]
MYO1A rs208210464 NC_037332.1 Exon region None Chinese cattle, Angus, Indian zebu T, RH, THI, and SR [24]
rs110123931 NC_037332.1 Exon region None Chinese cattle, Angus, Indian zebu T, RH, THI, and SR [24]
rs209999142 NC_037332.1 Exon region Phenylalanine to Serine Chinese cattle, Angus, Indian zebu T, RH, THI, and SR [24]
rs135771836 NC_037332.1 Intron region None Chinese cattle, Angus, Indian zebu T, RH, THI, and SR [24]
g.56390345 A>G NC_037332.1 Coding region Isoleucine into Valine Chinese cattle, Angus, Burma T, RH, THI, and SR [109]
ATP1A1 −14103G>A NC_007301.3 Exon 14 None Holstein cattle HTC and RR [25]
−14242C>T NC_007301.3 Intron 14 None Holstein cattle HTC and RR [25]
C2789A NC_007301.3 Exon 17 None Holstein cattle Heat resistance [110]
C2789A NC_007301.3 Exon 17 None Vrindavani and Tharparkar HTC and RT [112]
T27008243C Not mentioned Not mentioned Not mentioned Sahiwal RR [111]
A27008223G Not mentioned Not mentioned Not mentioned Karan Fries RR [111]

SNP, single-nucleotide polymorphism; HSP70, heat shock proteins 70; LA, lactate concentration in blood; SOD, superoxide dismutase concentration in blood; LPO, lipid perioxide concentration in blood; DA, dopamine concentration in blood; RT, rectal temperature; ART, average rectal temperature; HTC, heat tolerance coefficient; RR, respiration rate; ARR, average respiration rate; BT, body temperature; T, annual temperature; RH, relative humidity; THI, temperature humidity index; SR, average annual sunshine hours (100-cloudiness); PCE, potassium content in erythrocytes; HSF1, heat shock factors 1; EIF2AK4, eukaryotic translation initiation factor 2-alpha kinase 4; HSBP1, heat shock factor binding protein 1; MYO1A, myosin-1a; ATP1A1, ATPase Na+/K+ transporting subunit alpha 1.

REFERENCES

1. Chang-Fung-Martel J, Harrison MT, Brown JN, et al. Negative relationship between dry matter intake and the temperature-humidity index with increasing heat stress in cattle: a global meta-analysis. Int J Biometeorol 2021; 65:2099–109. https://doi.org/10.1007/s00484-021-02167-0
crossref pmid pmc
2. Corazzin M, Saccà E, Lippe G, et al. Effect of heat stress on dairy cow performance and on expression of protein metabolism genes in mammary cells. Animals 2020; 10:2124 https://doi.org/10.3390/ani10112124
crossref pmid pmc
3. Summer A, Lora I, Formaggioni P, Gottardo F. Impact of heat stress on milk and meat production. Anim Front 2019; 9:39–46. https://doi.org/10.1093/af/vfy026
crossref pmid pmc
4. Liu J, Li L, Chen X, Lu Y, Wang D. Effects of heat stress on body temperature, milk production, and reproduction in dairy cows: a novel idea for monitoring and evaluation of heat stress — A review. Asian-Australas J Anim Sci 2019; 32:1332–9. https://doi.org/10.5713/ajas.18.0743
crossref pmid pmc
5. Monteiro APA, Tao S, Thompson IMT, Dahl GE. In utero heat stress decreases calf survival and performance through the first lactation. J Dairy Sci 2016; 99:8443–50. https://doi.org/10.3168/jds.2016-11072
crossref pmid
6. Laporta J, Fabris TF, Skibiel AL, et al. In utero exposure to heat stress during late gestation has prolonged effects on the activity patterns and growth of dairy calves. J Dairy Sci 2017; 100:2976–84. https://doi.org/10.3168/jds.2016-11993
crossref pmid
7. Boonkum W, Misztal I, Duangjinda M, Pattarajinda V, Tumwasorn S, Buaban S. Short communication: genetic effects of heat stress on days open for Thai Holstein crossbreds. J Dairy Sci 2011; 94:1592–6. https://doi.org/10.3168/jds.2010-3491
crossref pmid
8. El-Tarabany MS, El-Bayoumi KM. Reproductive performance of backcross Holstein × Brown Swiss and their Holstein contemporaries under subtropical environmental conditions. Theriogenology 2015; 83:444–8. https://doi.org/10.1016/j.theriogenology.2014.10.010
crossref pmid
9. St-Pierre NR, Cobanov B, Schnitkey G. Economic losses from heat stress by US livestock industries. J Dairy Sci 2003; 86:E52–E77. https://doi.org/10.3168/jds.S0022-0302(03)74040-5
crossref
10. Johnson JS. Heat stress: Impact on livestock well-being and productivity and mitigation strategies to alleviate the negative effects. Anim Prod Sci 2018; 58:1404–13. https://doi.org/10.1071/AN17725
crossref
11. Fournel S, Ouellet V, Charbonneau É. Practices for alleviating heat stress of dairy cows in humid continental climates: a literature review. Animals 2017; 7:37 https://doi.org/10.3390/ani7050037
crossref pmid pmc
12. Augenstein SM, Harrison MA, Klopatek SC, Oltjen JW. Heat stress alleviation and dynamic temperature measurement for growing beef cattle. Transl Anim Sci 2020; 4:S178–81. https://doi.org/10.1093/tas/txaa144
crossref pmid pmc
13. Ansari-Mahyari S, Ojali MR, Forutan M, Riasi A, Brito LF. Investigating the genetic architecture of conception and non-return rates in Holstein cattle under heat stress conditions. Trop Anim Health Prod 2019; 51:1847–53. https://doi.org/10.1007/s11250-019-01875-5
crossref pmid
14. Carabaño MJ. The challenge of genetic selection for heat tolerance: the dairy cattle example. Adv Anim Biosci 2016; 7:218–222. https://doi.org/10.1017/s2040470016000169
crossref
15. Santana ML, Bignardi AB, Pereira RJ, Stefani G, El Faro L. Genetics of heat tolerance for milk yield and quality in Holsteins. Animal 2017; 11:4–14. https://doi.org/10.1017/S1751731116001725
crossref pmid
16. Beuzen ND, Stear MJ, Chang KC. Molecular markers and their use in animal breeding. Vet J 2000; 160:42–52. https://doi.org/10.1053/tvjl.2000.0468
crossref pmid
17. Abbas Z, Hu L, Fang H, et al. Association analysis of polymorphisms in the 5′ flanking region of the hsp70 gene with blood biochemical parameters of lactating holstein cows under heat and cold stress. Animals 2020; 10:2016 https://doi.org/10.3390/ani10112016
crossref pmid pmc
18. Onasanya GO, Msalya GM, Thiruvenkadan AK, et al. Single nucleotide polymorphisms at heat shock protein 90 gene and their association with thermo-tolerance potential in selected indigenous Nigerian cattle. Trop Anim Health Prod 2020; 52:1961–70. https://doi.org/10.1007/s11250-020-02222-9
crossref pmid
19. Rong Y, Zeng M, Guan X, et al. Association of HSF1 genetic variation with heat tolerance in chinese cattle. Animals 2019; 9:1027 https://doi.org/10.3390/ani9121027
crossref pmid pmc
20. Wang K, Cao Y, Rong Y, et al. A novel SNP in EIF2AK4 gene is associated with thermal tolerance traits in Chinese cattle. Animals 2019; 9:375 https://doi.org/10.3390/ani9060375
crossref pmid pmc
21. Wang Y, Huang J, Xia P, et al. Genetic variations of HSBP1 gene and its effect on thermal performance traits in Chinese Holstein cattle. Mol Biol Rep 2013; 40:3877–82. https://doi.org/10.1007/s11033-012-1977-1
crossref pmid
22. Verma N, Gupta ID, Verma A, et al. Novel SNPs in HSPB8 gene and their association with heat tolerance traits in Sahiwal indigenous cattle. Trop Anim Health Prod 2016; 48:175–80. https://doi.org/10.1007/s11250-015-0938-9
crossref pmid
23. Zeng L, Cao Y, Wu Z, et al. A missense mutation of the HSPB7 gene associated with heat tolerance in Chinese indicine cattle. Animals 2019; 9:554 https://doi.org/10.3390/ani9080554
crossref pmid pmc
24. Jia P, Cai C, Qu K, et al. Four novel SNPs of MYO1A gene associated with heat-tolerance in Chinese cattle. Animals 2019; 9:964 https://doi.org/10.3390/ani9110964
crossref pmid pmc
25. Liu YX, Zhou X, Li DQ, Cui QW, Wang GL. Association of ATP1A1 gene polymorphism with heat tolerance traits in dairy cattle. Genet Mol Res 2010; 9:891–6. https://doi.org/10.4238/vol9-2gmr769
crossref pmid
26. Garbuz DG. Regulation of heat shock gene expression in response to stress. Mol Biol 2017; 51:352–67. https://doi.org/10.1134/S0026893317020108
crossref
27. Kumar R, Gupta ID, Verma A, et al. Novel SNP identification in exon 3 of HSP90AA1 gene and their association with heat tolerance traits in Karan Fries (Bos taurus × Bos indicus) cows under tropical climatic condition. Trop Anim Health Prod 2016; 48:735–40. https://doi.org/10.1007/s11250-016-1016-7
crossref pmid
28. Koltes JE, Koltes DA, Mote BE, Tucker J, Hubbell DS. Automated collection of heat stress data in livestock: new technologies and opportunities. Transl Anim Sci 2018; 2:319–23. https://doi.org/10.1093/tas/txy061
crossref pmid pmc
29. Deb R, Fonsêca VDFC, Payan-Carreira R, Sejian V, Lees AM. Editorial: Genetic basis of thermoregulation in livestock. Front Vet Sci 2022; 9:839612 https://doi.org/10.3389/fvets.2022.839612
crossref pmid pmc
30. Carabaño MJ, Ramón M, Menéndez-Buxadera A, Molina A, Díaz C. Selecting for heat tolerance. Anim Front 2019; 9:62–8. https://doi.org/10.1093/af/vfy033
crossref pmid pmc
31. Dikmen S, Cole JB, Null DJ, Hansen PJ. Heritability of rectal temperature and genetic correlations with production and reproduction traits in dairy cattle. J Dairy Sci 2012; 95:3401–5. https://doi.org/10.3168/jds.2011-4306
crossref pmid
32. Nguyen TTT, Bowman PJ, Haile-Mariam M, Pryce JE, Hayes BJ. Genomic selection for tolerance to heat stress in Australian dairy cattle. J Dairy Sci 2016; 99:2849–62. https://doi.org/10.3168/jds.2015-9685
crossref pmid
33. Feder ME, Hofmann GE. Heat-shock proteins, molecular chaperones, and the stress response: evolutionary and ecological physiology. Annu Rev Physiol 1999; 61:243–82. https://doi.org/10.1146/annurev.physiol.61.1.243
crossref pmid
34. Chirico WJ, Waters MG, Blobel G. 70K heat shock related proteins stimulate protein translocation into microsomes. Nature 1988; 332:805–10. https://doi.org/10.1038/332805a0
crossref pmid
35. Hassan FU, Nawaz A, Rehman MS, Alid MA, Dilshad SMR, Yang C. Prospects of HSP70 as a genetic marker for thermo-tolerance and immuno-modulation in animals under climate change scenario. Anim Nutr (Zhongguo Xu Mu Shou Yi Xue Hui) 2019; 5:340–50. https://doi.org/10.1016/j.aninu.2019.06.005
crossref pmid pmc
36. Kishore A, Sodhi M, Kumari P, et al. Peripheral blood mononuclear cells: a potential cellular system to understand differential heat shock response across native cattle (Bos indicus), exotic cattle (Bos taurus), and riverine buffaloes (Bubalus bubalis) of India. Cell Stress Chaperones 2014; 19:613–21. https://doi.org/10.1007/s12192-013-0486-z
crossref pmid pmc
37. Gade N, Mahapatra RK, Sonawane A, Singh VK, Doreswamy R, Saini M. Molecular characterization of heat shock protein 70-1 gene of goat (Capra hircus). Mol Biol Int 2010; 2010:108429 https://doi.org/10.4061/2010/108429
crossref pmid pmc
38. Gallagher DSJ, Grosz MD, Womack JE, Skow LC. Chromosomal localization of HSP70 genes in cattle. Mamm Genome 1993; 4:388–90. https://doi.org/10.1007/BF00360590
crossref pmid
39. Dangi SS, Gupta M, Maurya D, et al. Expression profile of HSP genes during different seasons in goats (Capra hircus). Trop Anim Health Prod 2012; 44:1905–12. https://doi.org/10.1007/s11250-012-0155-8
crossref pmid
40. Kregel KC. Heat shock proteins: modifying factors in physiological stress responses and acquired thermotolerance. J Appl Physiol 2002; 92:2177–86. https://doi.org/10.1152/japplphysiol.01267.2001
crossref pmid
41. Pryce JE, Nguyen TTT, Cheruiyot EK, Marett L, Garner JB, Haile-Mariam M. Impact of hot weather on animal performance and genetic strategies to minimise the effect. Anim Prod Sci 2022; 62:726–35. https://doi.org/10.1071/AN21259
crossref
42. Xu Q, Wang YC, Liu R, et al. Differential gene expression in the peripheral blood of Chinese Sanhe cattle exposed to severe cold stress. Genet Mol Res. 2017. 16:gmr16029593 https://doi.org/10.4238/gmr16029593
crossref
43. Hu L, Ma Y, Liu L, et al. Detection of functional polymorphisms in the hsp70 gene and association with cold stress response in Inner-Mongolia Sanhe cattle. Cell Stress Chaperones 2019; 24:409–18. https://doi.org/10.1007/s12192-019-00973-5
crossref pmid pmc
44. Bhat S, Kumar P, Kashyap N, et al. Effect of heat shock protein 70 polymorphism on thermotolerance in Tharparkar cattle. Vet World 2016; 9:113–7. https://doi.org/10.14202/vetworld.2016.113-117
crossref pmid pmc
45. Onasanya GO, Msalya GM, Thiruvenkadan AK, et al. Heterozygous single-nucleotide polymorphism genotypes at heat shock protein 70 gene potentially influence thermo-tolerance among four Zebu breeds of Nigeria. Front Genet 2021; 12:642213 https://doi.org/10.3389/fgene.2021.642213
crossref pmid pmc
46. Prihandini PW, Primasari A, Aryogi A, Luthfi M, Hariyono DNH. Genetic polymorphisms of the 5’ untranslated regions of the HSP70 gene in Indonesian cattle populations. Vet World 2022; 15:168–72. https://doi.org/10.14202/vetworld.2022.168-172
crossref pmid pmc
47. Starkey L, Looper M, Banks A, Reiter S, Rosenkrans C. Identification of polymorphisms in the promoter region of the bovine heat shock protein gene and associations with bull calf weaning weight. J Anim Sci 2007; 85:Sppl 244

48. Rosenkrans CJ, Banks A, Reiter S, Looper M. Calving traits of crossbred Brahman cows are associated with Heat Shock Protein 70 genetic polymorphisms. Anim Reprod Sci 2010; 119:178–82. https://doi.org/10.1016/j.anireprosci.2010.02.005
crossref pmid
49. Mkize LS, Zishiri OT. Novel single nucleotide polymorphisms in the heat shock protein 70.1 gene in South African Nguni crossbred cattle. Trop Anim Health Prod 2020; 52:893–901. https://doi.org/10.1007/s11250-019-02088-6
crossref pmid
50. Basiricò L, Morera P, Primi V, Lacetera N, Nardone A, Bernabucci U. Cellular thermotolerance is associated with heat shock protein 70.1 genetic polymorphisms in Holstein lactating cows. Cell Stress Chaperones 2011; 16:441–8. https://doi.org/10.1007/s12192-011-0257-7
crossref pmid pmc
51. Deb R, Sajjanar B, Singh U, et al. Promoter variants at AP2 box region of Hsp70.1 affect thermal stress response and milk production traits in Frieswal cross bred cattle. Gene 2013; 532:230–5. https://doi.org/10.1016/j.gene.2013.09.037
crossref pmid
52. Öner Y, Keskin A, Üstüner H, Soysal D, Karakas V. Genetic diversity of the 3′ and 5′ untranslated regions of the HSP70.1 gene between native Turkish and Holstein Friesian cattle breeds. S Afr J Anim Sci 2017; 47:424–39. https://doi.org/10.4314/sajas.v47i2
crossref
53. Badri T, Alsiddig M, Lian L, Cai Y, Wang G. Single nucleotide polymorphisms in HSP70–1 gene associated with cellular heat tolerance in Chinese Holstein cows. Anim Gene 2021; 20:200114 https://doi.org/10.1016/j.angen.2021.200114
crossref
54. Pratt WB, Toft DO. Regulation of signaling protein function and trafficking by the hsp90/hsp70-based chaperone machinery. Exp Biol Med (Maywood) 2003; 228:111–33. https://doi.org/10.1177/153537020322800201
crossref pmid
55. Genest O, Wickner S, Doyle SM. Hsp90 and Hsp70 chaperones: Collaborators in protein remodeling. J Biol Chem 2019; 294:2109–20. https://doi.org/10.1074/jbc.REV118.002806
crossref pmid pmc
56. Mosser DD, Caron AW, Bourget L, Denis-Larose C, Massie B. Role of the human heat shock protein hsp70 in protection against stress-induced apoptosis. Mol Cell Biol 1997; 17:5317–27. https://doi.org/10.1128/MCB.17.9.5317
crossref pmid pmc
57. Wang P, Wang Y-F, Wang H, et al. HSP90 expression correlation with the freezing resistance of bull sperm. Zygote 2014; 22:239–45. https://doi.org/10.1017/S096719941300004X
crossref pmid
58. Zhang X-G, Hu S, Han C, et al. Association of heat shock protein 90 with motility of post-thawed sperm in bulls. Cryobiology 2015; 70:164–9. https://doi.org/10.1016/j.cryobiol.2014.12.010
crossref pmid
59. Chen B, Zhong D, Monteiro A. Comparative genomics and evolution of the HSP90 family of genes across all kingdoms of organisms. BMC Genomics 2006; 7:156 https://doi.org/10.1186/1471-2164-7-156
crossref pmid pmc
60. Csermely P, Schnaider T, Soti C, Prohászka Z, Nardai G. The 90-kDa molecular chaperone family: structure, function, and clinical applications. A comprehensive review. Pharmacol Ther 1998; 79:129–68. https://doi.org/10.1016/s0163-7258(98)00013-8
crossref pmid
61. Charoensook R, Gatphayak K, Sharifi AR, et al. Polymorphisms in the bovine HSP90AB1 gene are associated with heat tolerance in Thai indigenous cattle. Trop Anim Health Prod 2012; 44:921–8. https://doi.org/10.1007/s11250-011-9989-8
crossref pmid pmc
62. Sajjanar B, Deb R, Singh U, et al. Identification of SNP in HSP90AB1 and its association with the relative thermotolerance and milk production traits in Indian dairy cattle. Anim Biotechnol 2015; 26:45–50. https://doi.org/10.1080/10495398.2014.882846
crossref pmid
63. Badri TM, Chen KL, Alsiddig MA, Li L, Cai Y, Wang GL. Genetic polymorphism in Hsp90AA1 gene is associated with the thermotolerance in Chinese Holstein cows. Cell Stress Chaperones 2018; 23:639–51. https://doi.org/10.1007/s12192-017-0873-y
crossref pmid pmc
64. Gill JK, Arora JS, Sunil Kumar BV, Mukhopadhyay CS, Kaur S, Kashyap N. Cellular thermotolerance is independent of HSF 1 expression in zebu and crossbred non-lactating cattle. Int J Biometeorol 2017; 61:1687–93. https://doi.org/10.1007/s00484-017-1350-0
crossref pmid
65. Carpenter RL, Gökmen-Polar Y. HSF1 as a cancer biomarker and therapeutic target. Curr Cancer Drug Targets 2019; 19:515–24. https://doi.org/10.2174/1568009618666181018162117
crossref pmid pmc
66. Ritossa F. A new puffing pattern induced by temperature shock and DNP in drosophila. Experientia 1962; 18:571–3. https://doi.org/10.1007/BF02172188
crossref
67. Lindquist S. The heat-shock response. Annu Rev Biochem 1986; 55:1151–91. https://doi.org/10.1146/annurev.bi.55.070186.005443
crossref pmid
68. Sistonen L, Sarge KD, Morimoto RI. Human heat shock factors 1 and 2 are differentially activated and can synergistically induce hsp70 gene transcription. Mol Cell Biol 1994; 14:2087–99. https://doi.org/10.1128/mcb.14.3.2087-2099.1994
crossref pmid pmc
69. Edea Z, Dadi H, Dessie T, et al. Genome-wide scan reveals divergent selection among taurine and zebu cattle populations from different regions. Anim Genet 2018; 49:550–63. https://doi.org/10.1111/age.12724
crossref pmid
70. de Fátima Bretanha Rocha R, Baena MM, de Cássia Estopa A, et al. Differential expression of HSF1 and HSPA6 genes and physiological responses in Angus and Simmental cattle breeds. J Therm Biol 2019; 84:92–8. https://doi.org/10.1016/j.jtherbio.2019.06.002
crossref pmid
71. Li Q-L, Ju ZH, Huang JM, et al. Two novel SNPs in HSF1 gene are associated with thermal tolerance traits in Chinese Holstein cattle. DNA Cell Biol 2011; 30:247–54. https://doi.org/10.1089/dna.2010.1133
crossref pmid
72. Baena MM, Tizioto PC, Meirelles SLC, Regitano LC, de A. HSF1 and HSPA6 as functional candidate genes associated with heat tolerance in Angus cattle. Rev Bras Zootec. 2018. 47:e20160390 https://doi.org/10.1590/rbz4720160390
crossref
73. Sharma P, Sharma A, Sodhi M, et al. Characterizing binding sites of heat responsive microRNAs and their expression pattern in heat stressed PBMCs of native cattle, exotic cattle and riverine buffaloes. Mol Biol Rep 2019; 46:6513–24. https://doi.org/10.1007/s11033-019-05097-8
crossref pmid
74. Li QL, Zhang ZF, Xia P, et al. A SNP in the 3’-UTR of HSF1 in dairy cattle affects binding of target bta-miR-484. Genet Mol Res 2015; 14:12746–55. https://doi.org/10.4238/2015.October.19.18
crossref pmid
75. Morimoto H, Baba R. Cellular stress and eIF-2alpha kinase. J UOEH 2012; 34:331–8. https://doi.org/10.7888/juoeh.34.331
crossref pmid
76. Krishna KH, Kumar MS. Molecular evolution and functional divergence of eukaryotic translation initiation factor 2-alpha kinases. PLoS One 2018; 13:e0194335 https://doi.org/10.1371/journal.pone.0194335
crossref pmid pmc
77. Harding HP, Novoa I, Zhang Y, et al. Regulated translation initiation controls stress-induced gene expression in mammalian cells. Mol Cell 2000; 6:1099–108. https://doi.org/10.1016/s1097-2765(00)00108-8
crossref pmid
78. Roobol A, Roobol J, Bastide A, et al. p58IPK is an inhibitor of the eIF2α kinase GCN2 and its localization and expression underpin protein synthesis and ER processing capacity. Biochem J 2015; 465:213–25. https://doi.org/10.1042/BJ20140852
crossref pmid
79. Taniuchi S, Miyake M, Tsugawa K, Oyadomari M, Oyadomari S. Integrated stress response of vertebrates is regulated by four eIF2α kinases. Sci Rep 2016; 6:32886 https://doi.org/10.1038/srep32886
crossref pmid pmc
80. Donnelly N, Gorman AM, Gupta S, Samali A. The eIF2α kinases: their structures and functions. Cell Mol Life Sci 2013; 70:3493–511. https://doi.org/10.1007/s00018-012-1252-6
crossref pmid
81. Houston BJ, Nixon B, Martin JH, et al. Heat exposure induces oxidative stress and DNA damage in the male germ line. Biol Reprod 2018; 98:593–606. https://doi.org/10.1093/biolre/ioy009
crossref pmid
82. Liang L, Ma G, Chen K, et al. EIF2AK4 mutation in pulmonary veno-occlusive disease: A case report and review of the literature. Medicine (Baltimore) 2016; 95:e5030 https://doi.org/10.1097/MD.0000000000005030
crossref pmid pmc
83. Deng J, Harding HP, Raught B, et al. Activation of GCN2 in UV-irradiated cells inhibits translation. Curr Biol 2002; 12:1279–86. https://doi.org/10.1016/s0960-9822(02)01037-0
crossref pmid
84. Jiang HY, Wek RC. GCN2 phosphorylation of eIF2alpha activates NF-kappaB in response to UV irradiation. Biochem J 2005; 385:371–380. https://doi.org/10.1042/BJ20041164
crossref pmid pmc
85. Masson GR. Towards a model of GCN2 activation. Biochem Soc Trans 2019; 47:1481–8. https://doi.org/10.1042/BST20190331
crossref pmid pmc
86. Satyal SH, Chen D, Fox SG, Kramer JM, Morimoto RI. Negative regulation of the heat shock transcriptional response by HSBP1. Genes Dev 1998; 12:1962–74. https://doi.org/10.1101/gad.12.13.1962
crossref pmid pmc
87. Eroglu B, Min JN, Zhang Y, et al. An essential role for heat shock transcription factor binding protein 1 (HSBP1) during early embryonic development. Dev Biol 2014; 386:448–60. https://doi.org/10.1016/j.ydbio.2013.12.038
crossref pmid pmc
88. Liu X, Xu L, Liu Y, et al. Crystal structure of the hexamer of human heat shock factor binding protein 1. Proteins 2009; 75:1–11. https://doi.org/10.1002/prot.22216
crossref pmid
89. Chowdary TK, Raman B, Ramakrishna T, Rao CM. Mammalian Hsp22 is a heat-inducible small heat-shock protein with chaperone-like activity. Biochem J 2004; 381:379–87. https://doi.org/10.1042/BJ20031958
crossref pmid pmc
90. Dubińska-Magiera M, Niedbalska-Tarnowska J, Migocka-Patrzałek M, Posyniak E, Daczewska M. Characterization of Hspb8 in Zebrafish. Cells 2020; 9:1562 https://doi.org/10.3390/cells9061562
crossref pmid pmc
91. Chowdary TK, Raman B, Ramakrishna T, Rao CM. Interaction of mammalian Hsp22 with lipid membranes. Biochem J 2007; 401:437–45. https://doi.org/10.1042/BJ20061046
crossref pmid pmc
92. Shemetov AA, Seit-Nebi AS, Gusev NB. Structure, properties, and functions of the human small heat-shock protein HSP22 (HspB8, H11, E2IG1): a critical review. J Neurosci Res 2008; 86:264–9. https://doi.org/10.1002/jnr.21441
crossref pmid
93. Kuroyanagi G, Sakai G, Otsuka T, et al. HSP22 (HSPB8) positively regulates PGF2α-induced synthesis of interleukin-6 and vascular endothelial growth factor in osteoblasts. J Orthop Surg Res 2021; 16:72 https://doi.org/10.1186/s13018-021-02209-8
crossref pmid pmc
94. Fan H, Ding R, Liu W, et al. Heat shock protein 22 modulates NRF1/TFAM-dependent mitochondrial biogenesis and DRP1-sparked mitochondrial apoptosis through AMPK-PGC1α signaling pathway to alleviate the early brain injury of subarachnoid hemorrhage in rats. Redox Biol 2021; 40:101856 https://doi.org/10.1016/j.redox.2021.101856
crossref pmid pmc
95. Vos MJ, Kanon B, Kampinga HH. HSPB7 is a SC35 speckle resident small heat shock protein. Biochim Biophys Acta Mol Cell Res 2009; 1793:1343–53. https://doi.org/10.1016/j.bbamcr.2009.05.005
crossref
96. Ke L, Meijering RAM, Hoogstra-Berends F, et al. HSPB1, HSPB6, HSPB7 and HSPB8 protect against RhoA GTPase-induced remodeling in tachypaced atrial myocytes. PLoS One 2011; 6:e20395 https://doi.org/10.1371/journal.pone.0020395
crossref pmid pmc
97. Krief S, Faivre JF, Robert P, et al. Identification and characterization of cvHsp. A novel human small stress protein selectively expressed in cardiovascular and insulin-sensitive tissues. J Biol Chem 1999; 274:36592–600. https://doi.org/10.1074/jbc.274.51.36592
crossref pmid
98. Lin W, Yang Z, Lu Y, Zhao X. Refined purification of large amounts of rat cvHsp/HspB7 and partial biological characterization in vitro. Protein Pept Lett 2014; 21:503–10. https://doi.org/10.2174/092986652105140218121109
crossref pmid
99. Sun X, Li X, Jia H, et al. Effect of heat-shock protein B7 on oxidative stress in adipocytes from preruminant calves. J Dairy Sci 2019; 102:5673–85. https://doi.org/10.3168/jds.2018-15726
crossref pmid
100. Jin C, Shuai T, Tang Z. HSPB7 regulates osteogenic differentiation of human adipose derived stem cells via ERK signaling pathway. Stem Cell Res Ther 2020; 11:450 https://doi.org/10.1186/s13287-020-01965-4
crossref pmid pmc
101. Anzures-Olvera F, Véliz FG, de Santiago A, et al. The impact of hair coat color on physiological variables, reproductive performance and milk yield of Holstein cows in a hot environment. J Therm Biol 2019; 81:82–8. https://doi.org/10.1016/j.jtherbio.2019.02.020
crossref pmid
102. Katiyatiya CLF, Muchenje V. Hair coat characteristics and thermophysiological stress response of Nguni and Boran cows raised under hot environmental conditions. Int J Biometeorol 2017; 61:2183–94. https://doi.org/10.1007/s00484-017-1424-z
crossref pmid
103. Hayden SM, Wolenski JS, Mooseker MS. Binding of brush border myosin I to phospholipid vesicles. J Cell Biol 1990; 111:443–51. https://doi.org/10.1083/jcb.111.2.443
crossref pmid pmc
104. Berg JS, Powell BC, Cheney RE. A millennial myosin census. Mol Biol Cell 2001; 12:780–94. https://doi.org/10.1091/mbc.12.4.780
crossref pmid pmc
105. Munson S, Wang Y, Chang W, Bikle DD. Myosin 1a regulates osteoblast differentiation independent of intestinal calcium transport. J Endocr Soc 2019; 3:1993–2011. https://doi.org/10.1210/js.2019-00171
crossref pmid pmc
106. Tyska MJ, Nambiar R. Myosin-1a: A motor for microvillar membrane movement and mechanics. Commun Integr Biol 2010; 3:64–6. https://doi.org/10.4161/cib.3.1.10141
crossref pmid pmc
107. McConnell RE, Tyska MJ. Myosin-1a powers the sliding of apical membrane along microvillar actin bundles. J Cell Biol 2007; 177:671–81. https://doi.org/10.1083/jcb.200701144
crossref pmid pmc
108. Nambiar R, McConnell RE, Tyska MJ. Control of cell membrane tension by myosin-I. Proc Natl Acad Sci USA 2009; 106:11972–7. https://doi.org/10.1073/pnas.0901641106
crossref pmid pmc
109. Cao Y, Jia P, Wu Z, et al. A novel SNP of MYO1A gene associated with heat-tolerance in Chinese cattle. Anim Biotechnol. 2020. Nov. 4[Epub]. https://doi.org/10.1080/10495398.2020.1837147
crossref
110. Liu Y, Li D, Li H, Zhou X, Wang G. A novel SNP of the ATP1A1 gene is associated with heat tolerance traits in dairy cows. Mol Biol Rep 2011; 38:83–8. https://doi.org/10.1007/s11033-010-0080-8
crossref pmid
111. Das R, Gupta I, Verma A, et al. Genetic polymorphisms in ATP1A1 gene and their association with heat tolerance in Jersey crossbred cows. India J Dairy Sci 2015; 68:50–4.

112. Kashyap N, Kumar P, Deshmukh B, et al. Association of ATP1A1 gene polymorphism with thermotolerance in Tharparkar and Vrindavani cattle. Vet World 2015; 8:892–7. https://doi.org/10.14202/vetworld.2015.892-897
crossref pmid pmc
113. Zimin AV, Delcher AL, Florea L, et al. A whole-genome assembly of the domestic cow, Bos taurus. Genome Biol 2009; 10:R42 https://doi.org/10.1186/gb-2009-10-4-r42
crossref pmid pmc
114. Fujii N, Kenny GP, Amano T, et al. Na(+)-K(+)-ATPase plays a major role in mediating cutaneous thermal hyperemia achieved by local skin heating to 39°C. J Appl Physiol 2021; 131:1408–16. https://doi.org/10.1152/japplphysiol.00073.2021
crossref pmid
115. Vague P, Dufayet D, Coste T, Moriscot C, Jannot MF, Raccah D. Association of diabetic neuropathy with Na/K ATPase gene polymorphism. Diabetologia 1997; 40:506–11. https://doi.org/10.1007/s001250050708
crossref pmid
116. Rajamanickam GD, Kastelic JP, Thundathil JC. Na/K-ATPase regulates bovine sperm capacitation through raft- and non-raft-mediated signaling mechanisms. Mol Reprod Dev 2017; 84:1168–82. https://doi.org/10.1002/mrd.22879
crossref pmid
117. Blanco G. Na,K-ATPase subunit heterogeneity as a mechanism for tissue-specific ion regulation. Semin Nephrol 2005; 25:292–303. https://doi.org/10.1016/j.semnephrol.2005.03.004
crossref pmid
118. Elayadeth-Meethal M, Thazhathu Veettil A, Asaf M, et al. Comparative expression profiling and sequence characterization of ATP1A1 gene associated with heat tolerance in tropically adapted cattle. Animals 2021; 11:2368 https://doi.org/10.3390/ani11082368
crossref pmid pmc
119. Imran S, Khan MS, Hassan FU, Qureshi ZI. Genetic characterization of cholistani breed of cattle for atp1a1 gene and its association to heat tolerance traits. Pakistan J Agric Sci 2021; 58:229–34. https://doi.org/10.21162/PAKJAS/21.11
crossref
120. Barendse W, Reverter A, Bunch RJ, et al. A validated whole-genome association study of efficient food conversion in cattle. Genetics 2007; 176:1893–905. https://doi.org/10.1534/genetics.107.072637
crossref pmid pmc
121. Garner JB, Douglas ML, Williams SRO, et al. Genomic Selection improves heat tolerance in dairy cattle. Sci Rep 2016; 6:34114 https://doi.org/10.1038/srep34114
crossref pmid pmc
122. Zhou L, Ding X, Zhang Q, Wang Y, Lund MS, Su G. Consistency of linkage disequilibrium between Chinese and Nordic Holsteins and genomic prediction for Chinese Holsteins using a joint reference population. Genet Sel Evol 2013; 45:7 https://doi.org/10.1186/1297-9686-45-7
crossref pmid pmc
123. Lalrengpuii S, Gupta ID, Verma A, Das R, Chaudhari MV. Association of single nucleotide polymorphism of Hsp90ab1 gene with thermotolerance and milk yield in Sahiwal cows. African J Biochem Res 2015; 9:99–103. https://doi.org/10.5897/ajbr2015.0837
crossref


Editorial Office
Asian-Australasian Association of Animal Production Societies(AAAP)
Room 708 Sammo Sporex, 23, Sillim-ro 59-gil, Gwanak-gu, Seoul 08776, Korea   
TEL : +82-2-888-6558    FAX : +82-2-888-6559   
E-mail : editor@animbiosci.org               

Copyright © 2024 by Asian-Australasian Association of Animal Production Societies.

Developed in M2PI

Close layer
prev next