Next Article in Journal
Total and Differential Cell Counts as a Tool to Identify Intramammary Infections in Cows after Calving
Next Article in Special Issue
Research Progress on Oxidative Stress and Its Nutritional Regulation Strategies in Pigs
Previous Article in Journal
Researching Human-Cattle Interaction on Rangelands: Challenges and Potential Solutions
Previous Article in Special Issue
Modulation of Heat-Shock Proteins Mediates Chicken Cell Survival against Thermal Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Impacts of Heat Stress-Induced Oxidative Stress on the Milk Protein Biosynthesis of Dairy Cows

1
State Key Laboratory of Animal Nutrition, Institute of Animal Sciences, Chinese Academy of Agricultural Sciences, Beijing 100193, China
2
College of Animal Science and Technology, Yangzhou University, Yangzhou 225009, China
3
Joint Laboratory on Integrated Crop-Tree-Livestock Systems of the Chinese Academy of Agricultural Sciences (CAAS), Ethiopian Institute of Agricultural Research (EIAR) and World Agroforestry Center (ICRAF), Beijing 100193, China
*
Authors to whom correspondence should be addressed.
Submission received: 1 February 2021 / Revised: 28 February 2021 / Accepted: 4 March 2021 / Published: 7 March 2021

Abstract

:

Simple Summary

Heat stress (HS) of dairy cows affects milk protein synthesis partially due to a decline in appetite and dry matter intake (DMI). Several published studies indicate that HS causes oxidative stress (OS) with high levels of free radicals in tissues. Furthermore, the involvement of reactive oxygen species (ROS) in inducing apoptosis, in reducing mammary epithelial cell numbers, and in endocrine disruptions are proposed as the main mechanisms by which HS-induced OS modifies milk protein synthesis. However, challenges remain in determining the levels of apoptosis in vivo, as well as the tracking of the sources of ROS formation. Therefore, further investigations are required.

Abstract

Heat stress (HS) is one of the most important factors posing harm to the economic wellbeing of dairy industries, as it reduces milk yield as well as milk protein content. Recent studies suggest that HS participates in the induction of tissue oxidative stress (OS), as elevated levels of reactive oxygen species (ROS) and mitochondrial dysfunction were observed in dairy cows exposed to hot conditions. The OS induced by HS likely contributes to the reduction in milk protein content, since insulin resistance and apoptosis are promoted by OS and are negatively associated with the synthesis of milk proteins. The apoptosis in the mammary gland directly decreases the amount of mammary epithelial cells, while the insulin resistance affects the regulation of insulin on mTOR pathways. To alleviate OS damages, strategies including antioxidants supplementation have been adopted, but caution needs to be applied as an inappropriate supplement with antioxidants can be harmful. Furthermore, the complete mechanisms by which HS induces OS and OS influences milk protein synthesis are still unclear and further investigation is needed.

1. Introduction

Exposure of farm animals to high summer environmental temperatures is globally accepted to negatively affect animal husbandry, and no livestock is so more affected than dairy cows. Generally, dairy cows suffer from heat stress (HS) once the heat accumulated surpasses the animals’ capacity to dissipate this heat [1]. HS can be defined as the symptom of nonspecific immune response exhibited by animals in hot conditions according to the stress theory proposed by Selye [2], and it negatively impacts the productive traits of livestock. For dairy cows, HS could damage their productive traits including milk quality and milk yield [3].In China, an analysis of eight major milk producing provinces indicated that hot conditions in summer were associated with substantially lower milk production of dairy cows than at other times of the year [4], which causes economic losses in the dairy industry. Furthermore, the production of milk protein, an important quality parameter, is reduced by HS. [5,6,7] In many studies, this reduction was mainly attributed to a decline in feed intake [8]. However, with the assistance of pair-feeding studies, recent HS research indicated that decreased feed intake can only partially account for the reduction in both milk production and milk protein yield [7,9,10], suggesting additional mechanisms may exist in HS to directly disturb milk protein synthesis.
Reactive oxygen species (ROS) include superoxide anions(O2), hydrogen peroxide (H2O2), free hydroxyl radicals (OH) [11], and other products of aerobic metabolism [12]. These metabolites have dual functions in body regulation. On the one hand, ROS are related to maintaining the immune systems, for instance ROS can assist immunocytes in removing pathogens [13], and some others, such as H2O2, can act as second messengers for intercellular information exchange through redox signals [14]. On the other hand, increased ROS aggravates metabolic dysfunctions and even cause cell death, as ROS can modify highly reactive cellular macromolecules including organism lipids, proteins, and deoxyribonucleic acid (DNA) [15,16]. To prevent oxidative damage, animals are equipped with a defense mechanism, namely antioxidants, to counterbalance the effects of ROS, and these important compounds can be divided into enzymatic as well as nonenzymatic antioxidants [17]. When ROS are produced faster than the neutralizing ability of antioxidants, metabolic dysfunctions finally cause oxidative stress (OS) [12,18].
The involvement of HS as an inducer of OS in dairy cows has been acknowledged [19,20,21]. The decreased content of antioxidants in Holstein cows exposed to hot conditions suggests that HS participates in enhancing the formation of ROS [22,23,24]. This is notable, considering that DMI being reduced by HS may help to counteract OS by decreasing the absorption of some compounds from the digestive system. Since it is reported that OS participates in mTOR pathway regulation, as well as mediating apoptosis in mammary epithelial cells [25,26,27], the imbalance between ROS and antioxidants caused by HS may contribute to the reduction in milk protein through these two mechanisms. However, only a few studies have focused on the role of OS under HS on milk synthesis to comprehensively explain the effect of HS on milk protein synthesis of dairy cows. Hence, the current review aims to illustrate how OS induced by HS takes part in this process.

2. Heat Stress Associates with the Oxidative Stress in Body

2.1. The Determination of Oxidative Status in HS Cows

ROS are metabolites produced in the process of mitochondrial electron transport chain reaction [28], and these reactions are mostly mediated through redox chemistry of metal ions. Generally, transition metals in the body except for copper (Cu) contain an electron in the outermost shell while copper can easily loss or gain electrons even though it has a full outer shell [29]. Thus, iron as well as copper in the body become the common catalysts to induce oxidation reaction, as the electron can be considered as a free radical. The reduced forms of metal ions participate in a Fenton reaction where HO˙ is produced, then the ions in oxidative forms can return back to reduced forms by the Haber-Weiss reaction and participate in the Fenton reaction again [30,31]. As a result, redox active metals are likely to lead to the formation of OH˙.
Fenton reaction:
Metal(n+1) + H2O2 → Metal(n+1)+ + HO˙ + H2O
Haber-Weiss reaction
Metal(n+1)+ + 2O2−˙/AA/GSH → Metal(n+1) + O2/AA˙/GSSG
  • AA = ascorbic acid
  • GSH = reduced glutathione
  • GSSG = Oxidized glutathione
Obtaining direct measurements of ROS in vivo are difficult since the ROS has a very short half-life [32]. A few techniques including electron spin resonance (ESR) can detect free radicals directly relying on unpaired electrons of free radicals [33]. Unpaired electrons of free radicals behave as magnet so they can align either in a parallel or antiparallel manner when exposed to external magnetic fields [34]. Therefore, the two energy levels free radicals created can be varied and absorbed. Hence, free radicals can be identified by the absorption spectrum obtained via ESR [29]. However, the steady-state requirement limits the use of ESR in vivo; even though radicals can be accumulated to the level that does permit detection by ESR and spin trapping, the metabolism of the trap as well as the delicate nature of the trap itself may produce radicals in its absence [35]. Thus, a combination of ESR and spin trapping should be carefully used in vivo.
Indirect methods are widely selected for these tasks because the measuring effects caused by ROS instead of the total amount of generated ROS are easier to use. The techniques including fingerprinting methods and the measurements of antioxidant defense systems have been developed to detect oxidative status [29]. Fingerprinting is an alternative to trapping; its aim is to measure and quantify products of damage caused by ROS such as proteins, lipids, and DNA, and these end-products should specifically consider oxidative damage [36,37]. The determination of an antioxidant defense system is mainly achieved by measuring the enzymes such as superoxide dismutase (SOD) and catalase; there are two approaches available for this assay: one is to measure all the individual antioxidants recognized, while the other approach is to measure the capacity or activity of antioxidants by subjecting them to a controlled OS condition [38]. However, the synergistic effects among antioxidants and the influence of unknown antioxidant substances causes measuring the individual antioxidants to be time consuming and technically demanding to implement. In fact, the measurement of all kind of ROS within cells or overall antioxidative status at one specific time can be quite difficult since none of the parameters can predict the development of diseases induced by prolonged OS [39]. Therefore, it is important to employ multiple detection assays and consider the measurement of different parameters to enhance a method’s validity.

2.2. Heat Stress Contributes to the Overproduction of ROS

Oxidative stress occurs when the balance between pro-oxidants and antioxidants is disturbed and leads to an elevation in concentrations of free radicals and ROS [40,41]. ROS are metabolites produced in the process of mitochondrial electron transport chain reaction [28] and the vast majority are related to an increase in cellular aerobic respiration [12,28]. Various oxidase pathways such as nicotinamide adenine dinucleotide phosphate oxidase (NADPH oxidase) can generate ROS [42,43] when NADPH is converted to NADP+. The ROS produced in mitochondria are mainly superoxide anions, which can go on to generate hydrogen peroxides under the catalysis of superoxide dismutase (SOD) in the inner membrane of the mitochondria [44].
Several studies have demonstrated that HS promotes OS and ROS production [23,45], while HS can activate NADPH oxidase and increase the ratio of NADP+/NADPH, and it has been shown that heat treatment can enhance this ratio without any changes to NADPH subunits or the NADPH oxidase 1 (NOX1) mRNA expression [46]. Furthermore, the knockdown of NOX1 significantly inhibited ROS production, suggesting that NOX1 plays an important role in the process of HS inducing ROS [46].
HS leads to an overproduction of transition metal ions (TMI) by increasing the rate of iron release from ferritin [47]. TMI can donate electron to oxygen, forming superoxide anion and hydrogen peroxide [48,49], while the electron capture by oxygen is also conductive to the formation of H2O2, and hydroxyl radicals via the Fenton reaction, as mentioned above [50].
Hydrogen peroxides are able to diffuse across cell membranes because of their small size and relatively benign reactivity, and they can go on to modify different proteins including phosphatases, transcription factors, ion channels, etc. [51]. While hydroxyl radicals have a very-short half-life, they can react with many molecules that are extremely reactive [52]. Hence, an uncontrolled elevated level of ROS concentrations may induce free radical-mediated chain reactions which indiscriminately target and damage proteins, lipids, polysaccharides, and DNA [53,54,55,56].

2.3. Heat Stress Causes the Oxidative Stress by Inducing Mitochondrial Dysfunction

Previous studies have shown that Holstein cows under heat exposure have a higher level of thiobarbituric acid-reactive species and malondialdehyde than cows not exposed to heat stress [57,58]. Since these two factors are the major products of polyunsaturated fatty acid (PUFA) peroxidation [59], HS may be strongly associated with lipoperoxidation of mitochondrial membranes, which is another important source of free radicals production [41]. Furthermore, studies in nonneuronal cells and isolated mitochondria demonstrated that the amounts of energy produced by mitochondria are disrupted by HS [60]. Hyperthermia increases the permeability of the mitochondrial inner membrane and impairs oxidative phosphorylation [61,62]. The damage reported in mitochondria isolated from the hearts of rats treated with hyperthermia included a reduction in ATP synthesis, and a greater tendency to open mitochondrial permeability transition pores [61].
The heat stress response of cells is a highly conserved cascade of protein activation and gene expression changes which are regulated by heat shock transcription factors and that participate in the defense of stress damage [63]. Heat shock transcription factors are commonly found in organisms and have extensive homology, and are separated and activated from heat shock proteins (HSPs) after being stimulated by ROS, viruses, hyperthermia, and other factors [64]. The complexes formed can bind to the promoter region of heat stress elements, thereby regulating the transcription of HSPs [65,66]. Heat shock proteins are a protein super family, for which the function of many members involves molecular chaperones that function to stabilize other proteins that have been damaged by cell stress. These molecules are in the weight range from less than 10 KD to more than 110 KD [67] and both of their cellular locations and specific functions are different [68]. For example, the HSP70 family includes HSC73, HSP72, GRP78, and GRP75 that are able to assist with degrading unstable proteins as well as regulating the activity of transcription factors [69].
As summarized above, mitochondria are thermosensitive [70] and as such the effects of HS are on their most basic function, namely respiration [71]. Mitochondrial respiration is defined as a continuous reaction system composed of a series of hydrogen transfer and electron transfer reactions that are arranged in a specific order [72]. Energy in the form of electron flow source from the respiratory chain is converted into a H+ gradient through the inner mitochondrial membrane [73]. Then this H+ gradient dissipates through the adenosine triphosphate (ATP) synthase complex producing ATP [74]. The amount of ATP in cells is regulated by both mitochondrial and glycolytic ATP synthesis. While HS can inhibit mitochondrial ATP synthesis and result in the dysfunction of the electron transport chain [75], it has also been shown that HS can inactivate the complex I of the respiratory chain without affecting other complexes [76]. Consequently, the rate of electron flow through the electron transport chain is slowed down under HS, resulting in a decreased oxygen uptake but a increased formation concentration of O2 [77], and all of these factors above are correlated with the OS.

3. Oxidative Stress Induced by Heat Stress Reduces Milk Protein Synthesis

3.1. Heat Stress Directly Affects the Synthesis of Milk Protein

Animals suffer an HS decrease in metabolic heat production to maintain homeothermy, often causing decreased productivity. For dairy cows, HS decreases milk yield as well as milk protein, and this is partly due to the shortage of synthesis precursors in mammary glands [7,24]. Amino acids and glucoses are precursors for the synthesis of casein (the main component of milk protein), while rumen microbial protein is the main source of amino acids entering mammary glands. Hence the reduction in milk protein is highly correlated with decreased DMI under HS. However, recent studies have shown that the role of DMI cannot totally explain the negative effects of HS on milk protein synthesis [9]. The liver and mammary glands act cooperatively for the provision and utilization of milk protein precursors, while research using isotope labeling showed that HS may improve the ability of tissues (apart from mammary glands) to utilize glucose instead of the ability of hepatic gluconeogenesis [78]. Studies showed that HS can affect the lipid and carbohydrates metabolisms in cattle [79]. Gao et al. [7] conducted a pair-feed trial and found that HS significantly reduced the total amino acid and the content of amino acids including threonine, serine, glycine, cysteine, isoleucine, lysine, and arginine in blood compared to thermal neutral levels. Eventually, nutrient redistribution induced by HS causes the reduction in supplements of precursors of casein in the mammary gland and results in decreased milk protein synthesis.
Furthermore, signaling pathways dominated by Janus kinase 2/Signal transducers and activators of transcription (JAK2/STAT5) as well as mechanistic target of rapamycin (mTOR) are involved in the regulation of milk protein synthesis [80]. Related studies have found that HS inhibited the metabolic activity in mammary gland through RNA seq combined with data independent acquisition (DIA), while the inhibited genes relative to amino acid and glucose transporter reduces the overall casein synthesis ability [81]. Therefore, HS directly inhibits the synthesis of milk protein and results in the reduction of its content in milk.

3.2. Heat Stress Mediates Mammary Epithelial Cells Apoptosis through Oxidative Stress

Apoptosis is a process of energy dependent programmed cells death characterized by the specific degradation of cellular DNA. This process is regulated by specific genes and is morphologically manifested as nuclear pyknosis, cell membrane foaming, and apoptotic body formation and ends up with cells break down without any obvious lysis phenomenon [82]. Apoptosis can be activated by either intrinsic or extrinsic pathways, both of which are mediated by the caspase family of cysteine proteases [83]. The extrinsic pathway is induced by the activation of death receptors on the cell membrane, including factor associated suicide (FAS) and tumor necrosis factor receptor 1 (TNFR1) [84]. Meanwhile, the intrinsic pathway involves the mitochondrial and a complex of factors including stress conditions, chemical damages, and pharmaceutical damages [85].
Heat Stress can directly activate the mitochondrial pathway to mediate apoptosis [86]. Hyperthermia significantly decreases cell viability as well as resulting in abnormal mitochondrial morphology, which is an early signal of apoptosis [87]. The pro-apoptotic molecules are transferred to mitochondria and trigger apoptosis through the B cell lymphocytic-leukemia proto-oncogene (Bcl-2), while the Bcl-2 family includes both anti-apoptotic members like Bcl-2 and Bcl-xL and pro-apoptotic members such as Bax and Bak [88]. Under HS conditions, there are two mechanisms affecting outer mitochondrial membrane permeabilization to cytochrome c [89]. The first one relies on the outer mitochondrial membrane pores being regulated by Bcl-2 family members, for example, Bax/Bak oligomeric pores and Bax (Bak) voltage dependent anion channel hybrids [90,91,92]. The second mechanism involves mitochondrial Ca2+ overload caused by hyperthermia, where the overload results in persistent opening of the mitochondrial permeability transition pores, resulting in the nonspecific rupture of the outer mitochondrial membranes [93,94]. The rupture causes the releasing of cytochrome c, which binds to apoptotic protease activating factor 1 (Apaf-1) and forms the apoptosome. Hence, the apoptosome activates the initiator caspase-9 and its downstream effectors caspases-3 and caspase-7 [95], which then induces the subsequent apoptosis process. Furthermore, the overload of ROS or mitochondrial Ca2+ seems to assist the release of apoptotic cytochrome c, as was reported in some previous studies [96,97].
Since cellular stress, including OS, induces mitochondria to generate ROS [98], this has serious consequences including causing oxidative damages for mitochondrial DNA (mtDNA) [99]. Elevated levels of superoxide anions and hydroxyl radicals are considered to not only be associated with mtDNA damages, but also serve a major role in cell apoptosis [100]. However, the mechanism of mtDNA damage mediating apoptotic signaling is still not properly understood [100]. Furthermore, the involvement of ROS in apoptosis is controversial. The generation of free radicals or depletion of antioxidants induce apoptosis [101], but the apoptosis process can also occur at low oxygen tension (as Bcl-2 can show in the absence of ROS) [102]. The glutathione redox pair is an index of OS, hence a reduction in glutathione (GSH)/oxidized glutathione (GSSG) is proposed as a regulator of enzyme activities [103]. But the glutathione redox pair cannot evaluate the participance of OS in apoptosis process, as the glutathione depletion in apoptosis is due to an increased efflux of the reduced form; hence, this type of loss is nonoxidative [104]. As mentioned before, the regulation of cytochrome c released by mitochondria is important in apoptosis induction. That mitochondrial permeability transition channels (MPT) are independent of free radicals can explain the results obtained in both anaerobiosis and in aerobiosis. Related reports indicated that OS can regulate cytochrome c to be released from mitochondria via the mitochondrial permeability transition channels [25], suggesting that OS is a cause instead of a consequence in the apoptotic process, at least under aerobic conditions.
The reduction in the amount of lactating mammary epithelial cells can affect the synthesis of milk protein. Results of studies in vitro have shown that hyperthermia as well as the elevated OS contribute to the apoptosis of lactating mammary epithelial cells [62,105]. However, due to the limitation of cell apoptosis detection technology, it is still not yet fully proven that HS and its induced OS can induce apoptosis of mammary epithelial cells in vivo.

3.3. Oxidative Stress Results in Insulin Resistance and Associates with mTOR Pathway Regulation of Milk Protein Synthesis

The synthesis of milk protein is regulated by hormones including prolactin (PRL), glucocorticoids, thyroxine, and insulin [106], while the process of all protein synthesis, especially translation, is mediated by the mammalian target of the rapamycin (mTOR) pathway [107]. The mTOR pathway in mammals is a highly conserved and plays an important role in responding to external environmental signals of metabolism, growth, proliferation, and survival. mTOR is involved in the formation of mTORc1 and mTORc2 and serves as the catalytic core of these two signal complexes [108]. mTORc1 can not only regulate the synthesis of ribosome proteins, but exhibits a higher sensitivity to rapamycin than mTORc2, which plays an important role in regulating cell growth and reproduction [109,110]. The substrates that participate in mTORc1 regulation include ribosomal protein S6 kinases (S6K1 and S6K2) and eukaryotic initiation factor-4E-binding proteins (4EBPs) [111,112]. Also, 4EBP can inhibit the activity of eukaryotic initiation factor 4E (eIF4E), while the latter one can bind to the cap structure of mRNA. Furthermore, the site of 4EBPS binds to eIF4E is also the binding site of eIF4E and eIF4G, which affects the formation of translation initiation complexes [112]. mTORc1 can phosphorylate sites of 4EBPs to dissociate them from eIF4E, thereby promoting translation [112]. S6K1,a significant member of the AGC protein kinase family, can phosphorylate eEF2 kinase which in turn influences the initiation and extension of mRNA translation [113]. eEF2 is the translation elongation factor in mammalian cells, which assists with promoting the transfer of peptidyl RNA from A to P during translation and the subsequent formation of peptides [114]. eEF2 kinase specifically catalyzes the phosphorylation of eEF2 to inactivate the eEF2. Moreover, mTOR phosphorylates eEF2 kinase to inactivate the phosphorylated eEF2 and induces the dephosphorylation of eEF2, so as to promote translation elongation [115].
Insulin participates in the regulation of the mTOR signal pathway, as it can activate upstream protein kinase B (PKB) and further act on mTOR [26,116]. Insulin enhances phosphorylation levels of S6K1 as well as 4EBP1, promoting the transcription of mTOR signaling pathways; insulin can also activate eIF2 to further decrease the inhibitory effect of the massive activation of PKB [117]. Consequently, insulin is beneficial to protein synthesis and translation. However, former studies showed that insulin alone has an insignificant effect on the mTOR pathway, as it accelerates the transcription of milk protein genes and protein translation when interacting with other nutrients [118]. For example, growth hormone (GH) and insulin can synergistically promote protein synthesis and this effect is more intensive than that of insulin or GH alone [119]. Furthermore, as 80 percent of the glucose in body can be utilized by mammary glands, a certain concentration of insulin will reduce the uptake of metabolic substrates in other tissues [120]. This kind of nutrient distribution provides more synthesis precursors to the mammary gland, hence promoting the synthesis of milk protein. Additionally, one mouse research study showed that insulin can directly stimulate the expression of 28 protein synthesis genes, including 4 types of casein [121].
While OS has negative effects on milk production as described above, OS has also been shown to induce the insulin resistance (IR), impaired insulin secretion, and diabetes [27]. IR is defined as a state whereby normal concentrations of insulin are unable to stimulate adequate biological responses in insulin-sensitive tissues [122]. The accumulation of reactive molecules leads to the activation of multiple serine kinase cascades and/or inhibition of PTPases [123,124]. The insulin signaling pathway provides potential targets of these activated kinases, including the insulin receptor and the family of insulin receptor substrate (IRS) [125]. The ascended serine phosphorylation of IRS-1 and -2 could reduce those extent of tyrosine phosphorylation, which is consistent with the attenuation of insulin activity [126,127]. A published study found that H2O2 caused an increase in serine phosphorylation of IRS-1 as well as -2 but also decreased IRS-1 in 3T3-L1 adipocytes [128]. The motivated serine/ threonine phosphorylation accelerates the degradation of IRS-1 and then results in IR. Former research of dairy cows found that the tissues of liver and mammary glands contained a cooperative mechanism for the distribution and utilization of precursors required by milk protein synthesis [78]. Insulin has a multitude of effects on various metabolic pathways in different insulin-sensitive tissues. Except for the promoting protein synthesis as mentioned above, insulin participates in the suppression of glycogenolysis while stimulating glycolysis in liver and skeletal muscle [129,130]. The decreased biological response of insulin caused by IR affects the stimulation of protein synthesis induced by insulin via the mTOR pathway. Furthermore, effects of insulin on glycolysis as well as glycogenolysis are inhibited. These inhibitions together with HS damages contribute to the negative energy balance, triggering the mobilization of amino acids in blood for deamination reaction to meet energy requirement of organs. Thus, this redistribution reduces the supply of precursors and end up decreasing milk protein synthesis.

4. Strategies for Alleviating Heat Stress-Induced Oxidative Stress

To alleviate the OS induced by HS, strategies to reduce OS by neutralizing the overproduction of ROS or inhibiting its formation should be considered first [131,132]. But methods to reduce HS directly seems to be more effective in reducing its induced OS by either enhancing heat losses or lowering condition temperatures. These attempts of environmental modifications can be classified as shading, air cooling, ventilation, and spraying according to the different realization approach, and are usually combined to adopt in different practices [133]. For example, physical cooling strategies of the combination of spraying and ventilation are commonly used in dairy barns, in which the roof insulation was equipped as well. However, except for using air condition to cool cows, most parts of strategies cannot alter air temperature nor relative humidity in barns, which increase the difficulty of assessing the alleviating effects [134]. Few methods have been systematically evaluated. For example, in spraying, selecting a flow rate of 1.3 L/min to spray cows was reported to have the highest efficiency, while excessive water did not further reduce HS [135,136]; in shading, the total heat load in open barns can be reduced by 30% or more with a well-designed roof [137]; in ventilation, the addition of low volume and high speed fans was proved to significantly decrease the rectal temperature as well as respiration rate of dairy cows [138]. But considering that the climate can vary greatly in different areas, cooling strategies should be properly configured for different conditions and be cautiously adopted.
Apart from alleviating heat stress of dairy cows using direct cooling methods (fans and sprays), antioxidants supplementation has been proven useful in HS situations to reduce the occurrence of udder infection as well as improve milk protein content [139]. Numerous antioxidants contribute to the defense mechanism with their own specific functions [140]. A variety of antioxidants including vitamins, carotenoids, polyphenolics, and trace elements can be added as dietary supplements [141,142]. Among feed-derived antioxidants, vitamin E and selenium (Se) are always considered as the primary factors which contribute to defending against OS [143,144]. Vitamin E is in a group of fat-soluble vitamins including four tocopherols and four tocotrienols [145], which are hypothesized to act catalytically, being efficiently reduced from their free radical forms to their native states [146]. Specifically, vitamin E can transfer hydrogen atoms (H) to free radicals, as the O-H bond in tocopherol located at 323 KJ/mol is about 10% weaker than most of phenols [147,148]. This weak bond allows vitamin E to contribute H to hydrogen superoxide and other free radicals, hence the ROS are neutralized and damage is minimized [147]. Through the oxidation-reduction reaction of hydrogen donors such as vitamin C, the tocopherol radicals that are thus generated are reused as tocopherols. Since vitamin E is fat-soluble, it can also be incorporated into cell membranes to protect cells from oxidative damage [149]. Compared with vitamin E, Se is considered to be a more important defense system. This trace element participates in the synthesis of selenoproteins and is related to the maintenance of redox balance [150]. Studies indicated that nineteen selenoproteins in cows are involved in antioxidant defenses while a total of 25 selenoproteins are identified to participate in body regulations [151]. Some of the 19 antioxidative selenoproteins are members of glutathione peroxidase (GSH-Px) and thioredoxin reductases (TrxR) [151]. GSH-Px has the function of reducing lipid peroxides to alcohols and reducing free hydrogen peroxide to water, while catalyzing the conversion of glutathione to its oxidized form [152]. TrxR is the unique enzyme catalyzing the reduction of thioredoxin [152], while the thioredoxin system induces the formation of reduced disulfide bonds in cells and can alleviate the OS originated from oxygen metabolism by taking electrons from NADPH [153]. Consequently, the supplementation of diets with antioxidants contributes to build up the defense system and accordingly alleviates the extent of OS.
The preceding paragraph detailed the inclusion of the non-enzymatic antioxidants in diets, but the supplementation of plant natural extracts including polyphenols, flavonoids, tannins, and gallic acid has been attracting more attention in recent years due to their high efficiency as well as low residuals [154,155,156]. Among those different types of extracts tannins are the most-studied compounds [157], as they have the ability to scavenge free radicals by donating electrons to make those structures more stable and less toxic [158]. Tannins are classified into two groups: hydrolysable and condensed tannins [159]. A number of condensed as well as hydrolysable tannins were evaluated and reported to have the effects on super-oxide radicals, hydroxyl radicals, and nitric oxide [160,161,162].
The inappropriate use of antioxidant supplements is possible and may lead to the antioxidative imbalance and result in carcinogenesis [163]. As mentioned, ROS has a dual effect in tissues, hence slight OS is sometimes beneficial for the organism. In this case, antioxidant supplementation is not as beneficial as expected, for it may reduce some ROS which act as signaling molecules in important pathways. Therefore, it is always essential to keep the balance between ROS and antioxidants, no matter which kinds of strategies are adopted.

5. Conclusions

Heat stress causes an overproduction of ROS and mitochondrial dysfunction, leading to oxidative stress in dairy cows. Current published studies strongly suggest that oxidative stress as a result of heat stress contributes to a reduction in milk protein. This is caused by an increase in apoptosis in mammary gland tissues that directly reduces the number of mammary epithelial cells, while elevated levels of free radicals also damage milk protein synthesis by regulating signaling pathways. The dietary supplementation of antioxidants has been adopted to alleviate OS including that induced by HS, and this is a strategy that has been proven to be useful in certain circumstances. However, current technology has not been able to establish the real extent of apoptosis in mammary epithelial cells in vivo nor the accurate source of free radical formation in all situations (e.g., aerobic or anaerobic conditions). Further advances are needed to fully understand the effects of HS-induced OS on milk protein reduction, which have the potential to facilitate new methods to reduce the effects of high environmental temperatures on poor milk production in dairy cows.

Author Contributions

Conceptualization, Z.G. and D.B.; Drafting, review and editing of the manuscripts, Z.G., S.G., J.O., L.M., and D.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially supported by the National Natural Science Foundation of China (31872383), the National Key Research and Development Program of China (2018YFD0501600), the Scientific Research Project for Major Achievements of The Agricultural Science and Technology Innovation Program (ASTIP) (CAAS-ZDXT2019004, ASTIP-IAS07, CAAS-XTCX2016011-01), and the Beijing Dairy Industry Innovation Team (BAIC06-2020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data used in this study are included in the main text.

Acknowledgments

We give special thanks to members who have contributed to the relevant projects.

Conflicts of Interest

All the authors declare no conflict of interest.

References

  1. Akbarian, A.; Michiels, J.; Degroote, J.; Majdeddin, M.; Golian, A.; De Smet, S. Association between heat stress and oxidative stress in poultry; mitochondrial dysfunction and dietary interventions with phytochemicals. J. Anim. Sci. Biotechnol. 2016, 7, 37. [Google Scholar] [CrossRef] [Green Version]
  2. Selye, H. Studies on adaptation. Endocrinology 1937, 21, 169–188. [Google Scholar] [CrossRef]
  3. Baumgard, L.; Rhoads, R. Ruminant Nutrition Symposium: Ruminant production and metabolic responses to heat stress. J. Anim. Sci. 2012, 90, 1855–1865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ranjitkar, S.; Bu, D.; Van Wijk, M.; Ma, Y.; Ma, L.; Zhao, L.; Shi, J.; Liu, C.; Xu, J. Will heat stress take its toll on milk production in China? Clim. Chang. 2020, 161, 637–652. [Google Scholar] [CrossRef]
  5. Rhoads, M.; Rhoads, R.; VanBaale, M.; Collier, R.J.; Sanders, S.; Weber, W.; Crooker, B.; Baumgard, L. Effects of heat stress and plane of nutrition on lactating Holstein cows: I. Production, metabolism, and aspects of circulating somatotropin. J. Dairy Sci. 2009, 92, 1986–1997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Wheelock, J.; Rhoads, R.; VanBaale, M.; Sanders, S.; Baumgard, L. Effects of heat stress on energetic metabolism in lactating Holstein cows. J. Dairy Sci. 2010, 93, 644–655. [Google Scholar] [CrossRef]
  7. Gao, S.; Guo, J.; Quan, S.; Nan, X.; Fernandez, M.S.; Baumgard, L.; Bu, D. The effects of heat stress on protein metabolism in lactating Holstein cows. J. Dairy Sci. 2017, 100, 5040–5049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Fuquay, J. Heat stress as it affects animal production. J. Anim. Sci. 1981, 52, 164–174. [Google Scholar] [CrossRef]
  9. Cowley, F.; Barber, D.; Houlihan, A.; Poppi, D. Immediate and residual effects of heat stress and restricted intake on milk protein and casein composition and energy metabolism. J. Dairy Sci. 2015, 98, 2356–2368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Baumgard, L.H.; Rhoads, R.P., Jr. Effects of heat stress on postabsorptive metabolism and energetics. Annu. Rev. Anim. Biosci. 2013, 1, 311–337. [Google Scholar] [CrossRef] [Green Version]
  11. Halliwell, B. Biochemistry of Oxidative Stress; Portland Press Ltd.: London, UK, 2007. [Google Scholar]
  12. Schieber, M.; Chandel, N.S. ROS function in redox signaling and oxidative stress. Curr. Biol. 2014, 24, R453–R462. [Google Scholar] [CrossRef] [Green Version]
  13. Riley, P. Free radicals in biology: Oxidative stress and the effects of ionizing radiation. Int. J. Radiat. Biol. 1994, 65, 27–33. [Google Scholar] [CrossRef]
  14. Vajdovich, P. Free radicals and antioxidants in inflammatory processes and ischemia-reperfusion injury. Vet. Clin. N. Am. Small Anim. Pract. 2008, 38, 31–123. [Google Scholar] [CrossRef]
  15. Davies, K.J. Oxidative stress: The paradox of aerobic life. Biochem. Soc. Symp. 1995, 61, 1–31. [Google Scholar] [PubMed]
  16. Halliwell, B.; Whiteman, M. Measuring reactive species and oxidative damage in vivo and in cell culture: How should you do it and what do the results mean? Br. J. Pharmacol. 2004, 142, 231–255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Birben, E.; Sahiner, U.M.; Sackesen, C.; Erzurum, S.; Kalayci, O. Oxidative stress and antioxidant defense. World Allergy Organ. J. 2012, 5, 9–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Cross, C.E.; Halliwell, B.; Borish, E.T.; Pryor, W.A.; Ames, B.N.; Saul, R.L.; McCord, J.M.; Harman, D. Oxygen radicals and human disease. Ann. Intern. Med. 1987, 107, 526–545. [Google Scholar] [CrossRef] [PubMed]
  19. Altan, Ö.; Pabuçcuoğlu, A.; Altan, A.; Konyalioğlu, S.; Bayraktar, H. Effect of heat stress on oxidative stress, lipid peroxidation and some stress parameters in broilers. Br. Poult. Sci. 2003, 44, 545–550. [Google Scholar] [CrossRef] [PubMed]
  20. Sahin, K.; Onderci, M.; Sahin, N.; Gursu, M.; Kucuk, O. Dietary vitamin C and folic acid supplementation ameliorates the detrimental effects of heat stress in Japanese quail. J. Nutr. 2003, 133, 1882–1886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Lin, H.; Decuypere, E.; Buyse, J. Acute heat stress induces oxidative stress in broiler chickens. Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 2006, 144, 11–17. [Google Scholar] [CrossRef]
  22. Flanagan, S.W.; Moseley, P.L.; Buettner, G.R. Increased flux of free radicals in cells subjected to hyperthermia: Detection by electron paramagnetic resonance spin trapping. FEBS Lett. 1998, 431, 285–286. [Google Scholar] [CrossRef] [Green Version]
  23. Bernabucci, U.; Ronchi, B.; Lacetera, N.; Nardone, A. Markers of oxidative status in plasma and erythrocytes of transition dairy cows during hot season. J. Dairy Sci. 2002, 85, 2173–2179. [Google Scholar] [CrossRef] [Green Version]
  24. Guo, J.; Gao, S.; Quan, S.; Zhang, Y.; Bu, D.; Wang, J. Blood amino acids profile responding to heat stress in dairy cows. Asian Australas. J. Anim. Sci. 2018, 31, 47. [Google Scholar] [CrossRef] [Green Version]
  25. Yang, J.C.; Cortopassi, G.A. Induction of the mitochondrial permeability transition causes release of the apoptogenic factor cytochrome c. Free Radic. Biol. Med. 1998, 24, 624–631. [Google Scholar] [CrossRef]
  26. Burgos, S.; Dai, M.; Cant, J. Nutrient availability and lactogenic hormones regulate mammary protein synthesis through the mammalian target of rapamycin signaling pathway. J. Dairy Sci. 2010, 93, 153–161. [Google Scholar] [CrossRef] [Green Version]
  27. Ceriello, A.; Motz, E. Is oxidative stress the pathogenic mechanism underlying insulin resistance, diabetes, and cardiovascular disease? The common soil hypothesis revisited. Arterioscler. Thromb. Vasc. Biol. 2004, 24, 816–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Kärkönen, A.; Kuchitsu, K. Reactive oxygen species in cell wall metabolism and development in plants. Phytochemistry 2015, 112, 22–32. [Google Scholar] [CrossRef]
  29. Poljšak, B.; Jamnik, P. Methodology for oxidative state detection in biological systems. In Handbook of Free Radicals: Formation, Types and Effects; Nova Science Publishers, Inc.: Hauppauge, NY, USA, 2010. [Google Scholar]
  30. Nordberg, J.; Arnér, E.S. Reactive oxygen species, antioxidants, and the mammalian thioredoxin system. Free Radic. Biol. Med. 2001, 31, 1287–1312. [Google Scholar] [CrossRef]
  31. Kehrer, J.P. The Haber–Weiss reaction and mechanisms of toxicity. Toxicology 2000, 149, 43–50. [Google Scholar] [CrossRef]
  32. Sivanandham, V. Free radicals in health and diseases-a mini review. PharmacologyOnline 2011, 1, 1062–1077. [Google Scholar]
  33. Halliwell, B.; Gutteridge, J.M. Free Radicals in Biology and Medicine; Oxford University Press: Oxford, UK, 2015. [Google Scholar]
  34. Cheeseman, K.; Slater, T. An introduction to free radical biochemistry. Br. Med. Bull. 1993, 49, 481–493. [Google Scholar] [CrossRef] [PubMed]
  35. Rice-Evans, C.A.; Diplock, A.T.; Symons, M.R. Techniques in free radical research. Lab. Tech. Biochem. Mol. Biol. 1991, 22, 1–278. [Google Scholar]
  36. Heinecke, J.W. Oxidative stress: New approaches to diagnosis and prognosis in atherosclerosis. Am. J. Cardiol. 2003, 91, 12–16. [Google Scholar] [CrossRef]
  37. Miwa, S.; Beckman, K.B.; Muller, F. Oxidative Stress in Aging: From Model Systems to Human Diseases; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2008. [Google Scholar]
  38. Poljsak, B.; Šuput, D.; Milisav, I. Achieving the balance between ROS and antioxidants: When to use the synthetic antioxidants. Oxidative Med. Cell. Longev. 2013, 2013, 956792. [Google Scholar] [CrossRef] [PubMed]
  39. Halliwell, B. The wanderings of a free radical. Free Radic. Biol. Med. 2009, 46, 531–542. [Google Scholar] [CrossRef]
  40. Ganaie, A.; Ghasura, R.; Mir, N.; Bumla, N.; Sankar, G.; Wani, S. Biochemical and Physiological Changes during Thermal Stress in Bovines: A Review. 2013. Available online: https://www.sid.ir/en/journal/ViewPaper.aspx?id=325800 (accessed on 6 March 2021).
  41. Belhadj Slimen, I.; Najar, T.; Ghram, A.; Abdrrabba, M. Heat stress effects on livestock: Molecular, cellular and metabolic aspects, a review. J. Anim. Physiol. Anim. Nutr. 2016, 100, 401–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Segal, A.W.; Abo, A. The biochemical basis of the NADPH oxidase of phagocytes. Trends Biochem. Sci. 1993, 18, 43–47. [Google Scholar] [CrossRef]
  43. Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M.T.; Mazur, M.; Telser, J. Free radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem. Cell Biol. 2007, 39, 44–84. [Google Scholar] [CrossRef]
  44. Halliwell, B.; Gutteridge, J.M.C. The importance of free radicals and catalytic metal ions in human diseases. Mol. Asp. Med. 1985, 8, 89–193. [Google Scholar] [CrossRef]
  45. Lord-Fontaine, S.; Averill-Bates, D.A. Heat shock inactivates cellular antioxidant defenses against hydrogen peroxide: Protection by glucose. Free Radic. Biol. Med. 2002, 32, 752–765. [Google Scholar] [CrossRef]
  46. Moon, E.J.; Sonveaux, P.; Porporato, P.E.; Danhier, P.; Gallez, B.; Batinic-Haberle, I.; Nien, Y.-C.; Schroeder, T.; Dewhirst, M.W. NADPH oxidase-mediated reactive oxygen species production activates hypoxia-inducible factor-1 (HIF-1) via the ERK pathway after hyperthermia treatment. Proc. Natl. Acad. Sci. USA 2010, 107, 20477–20482. [Google Scholar] [CrossRef] [Green Version]
  47. Freeman, M.L.; Spitz, D.R.; Meredith, M.J. Does heat shock enhance oxidative stress? Studies with ferrous and ferric iron. Radiat. Res. 1990, 124, 288–293. [Google Scholar] [CrossRef] [PubMed]
  48. Powers, R.H.; Stadnicka, A.; Kalbfleish, J.H.; Skibba, J.L. Involvement of xanthine oxidase in oxidative stress and iron release during hyperthermic rat liver perfusion. Cancer Res. 1992, 52, 1699–1703. [Google Scholar]
  49. Agarwal, A.; Prabakaran, S.A. Mechanism, measurement, and prevention of oxidative stress in male reproductive physiology. Int. J. Electron. Bus. 2005, 43, 963–974. [Google Scholar]
  50. Liochev, S.I.; Fridovich, I. Superoxide and iron: Partners in crime. IUBMB Life 1999, 48, 157–161. [Google Scholar] [CrossRef] [PubMed]
  51. Thomas, S.R.; Witting, P.K.; Drummond, G.R. Redox control of endothelial function and dysfunction: Molecular mechanisms and therapeutic opportunities. Antioxid. Redox Signal. 2008, 10, 1713–1766. [Google Scholar] [CrossRef]
  52. Kirkinezos, I.G.; Moraes, C.T. Reactive oxygen species and mitochondrial diseases. Semin. Cell Dev. Biol. 2001, 12, 449–457. [Google Scholar] [CrossRef]
  53. Stadtman, E.R.; Levine, R.L. Protein oxidation. Ann. N. Y. Acad. Sci. 2000, 899, 191–208. [Google Scholar] [CrossRef] [PubMed]
  54. Rubbo, H.; Radi, R.; Trujillo, M.; Telleri, R.; Kalyanaraman, B.; Barnes, S.; Kirk, M.; Freeman, B.A. Nitric oxide regulation of superoxide and peroxynitrite-dependent lipid peroxidation. Formation of novel nitrogen-containing oxidized lipid derivatives. J. Biol. Chem. 1994, 269, 26066–26075. [Google Scholar] [CrossRef]
  55. Kaur, H.; Halliwell, B. Evidence for nitric oxide-mediated oxidative damage in chronic inflammation Nitrotyrosine in serum and synovial fluid from rheumatoid patients. FEBS Lett. 1994, 350, 9–12. [Google Scholar] [CrossRef] [Green Version]
  56. LeDoux, S.P.; Driggers, W.J.; Hollensworth, B.S.; Wilson, G.L. Repair of alkylation and oxidative damage in mitochondrial DNA. Mutat. Res. DNA Repair 1999, 434, 149–159. [Google Scholar] [CrossRef]
  57. Chandra, G.; Aggarwal, A. Effect of DL--Tocopherol Acetate on Calving Induced Oxidative Stress in Periparturient Crossbred Cows During Summer and Winter Seasons. Indian J. Anim. Nutr. 2009, 26, 204–210. [Google Scholar]
  58. Aengwanich, W.; Kongbuntad, W.; Boonsorn, T. Effects of shade on physiological changes, oxidative stress, and total antioxidant power in Thai Brahman cattle. Int. J. Biometeorol. 2011, 55, 741–748. [Google Scholar] [CrossRef]
  59. Gavino, V.; Miller, J.; Ikharebha, S.; Milo, G.; Cornwell, D. Effect of polyunsaturated fatty acids and antioxidants on lipid peroxidation in tissue cultures. J. Lipid Res. 1981, 22, 763–769. [Google Scholar] [CrossRef]
  60. White, M.G.; Saleh, O.; Nonner, D.; Barrett, E.F.; Moraes, C.T.; Barrett, J.N. Mitochondrial dysfunction induced by heat stress in cultured rat CNS neurons. J. Neurophysiol. 2012, 108, 2203–2214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Qian, L.; Song, X.; Ren, H.; Gong, J.; Cheng, S. Mitochondrial mechanism of heat stress-induced injury in rat cardiomyocyte. Cell Stress Chaperones 2004, 9, 281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Willis, W.; Jackman, M.; Bizeau, M.; Pagliassotti, M.; Hazel, J. Hyperthermia impairs liver mitochondrial function in vitro. Am. J. Physiol.-Regul. Integr. Comp. Physiol. 2000, 278, R1240–R1246. [Google Scholar] [CrossRef] [Green Version]
  63. Ohama, N.; Kusakabe, K.; Mizoi, J.; Zhao, H.; Kidokoro, S.; Koizumi, S.; Takahashi, F.; Ishida, T.; Yanagisawa, S.; Shinozaki, K. The transcriptional cascade in the heat stress response of Arabidopsis is strictly regulated at the level of transcription factor expression. Plant Cell 2016, 28, 181–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Guo, J.; Wu, J.; Ji, Q.; Wang, C.; Luo, L.; Yuan, Y.; Wang, Y.; Wang, J. Genome-wide analysis of heat shock transcription factor families in rice and Arabidopsis. J. Genet. Genom. 2008, 35, 105–118. [Google Scholar] [CrossRef]
  65. Lee, J.H.; Hübel, A.; Schöffl, F. Derepression of the activity of genetically engineered heat shock factor causes constitutive synthesis of heat shock proteins and increased thermotolerance in transgenic Arabidopsis. Plant J. 1995, 8, 603–612. [Google Scholar] [CrossRef] [PubMed]
  66. Lindquist, S.; Craig, E.A. The heat-shock proteins. Annu. Rev. Genet. 1988, 22, 631–677. [Google Scholar] [CrossRef]
  67. Schlesinger, M.J. Heat shock proteins. J. Biol. Chem. 1990, 265, 12111–12114. [Google Scholar] [CrossRef]
  68. Rappa, F.; Farina, F.; Zummo, G.; David, S.; Campanella, C.; Carini, F.; Tomasello, G.; Damiani, P.; Cappello, F.; De Macario, E.C. HSP-molecular chaperones in cancer biogenesis and tumor therapy: An overview. Anticancer Res. 2012, 32, 5139–5150. [Google Scholar] [PubMed]
  69. Skidmore, R.; Gutierrez, J.A.; Guerriero, V., Jr.; Kregel, K. HSP70 induction during exercise and heat stress in rats: Role of internal temperature. Am. J. Physiol.-Regul. Integr. Comp. Physiol. 1995, 268, R92–R97. [Google Scholar] [CrossRef] [PubMed]
  70. Chou, M.; Chen, Y.-M.; Lin, C.-Y. Thermotolerance of isolated mitochondria associated with heat shock proteins. Plant Physiol. 1989, 89, 617–621. [Google Scholar] [CrossRef] [Green Version]
  71. Ozawa, T.; Tanaka, M.; Suzuki, H.; Nishikimi, M. Structure and function of mitochondria: Their organization and disorders. Brain Dev. 1987, 9, 76–81. [Google Scholar] [CrossRef]
  72. Chance, B.; Williams, G. The respiratory chain and oxidative phosphorylation. Adv. Enzymol. Relat. Areas Mol. Biol. 1956, 17, 65–134. [Google Scholar]
  73. Mitchell, P. Vectorial chemiosmotic processes. Annu. Rev. Biochem. 1977, 46, 996–1005. [Google Scholar] [CrossRef]
  74. Noji, H.; Yoshida, M. The rotary machine in the cell, ATP synthase. J. Biol. Chem. 2001, 276, 1665–1668. [Google Scholar] [CrossRef] [Green Version]
  75. Monti, E.; Supino, R.; Colleoni, M.; Costa, B.; Ravizza, R.; Gariboldi, M.B. Nitroxide TEMPOL impairs mitochondrial function and induces apoptosis in HL60 cells. J. Cell. Biochem. 2001, 82, 271–276. [Google Scholar] [CrossRef]
  76. Downs, C.A.; Heckathorn, S.A. The mitochondrial small heat-shock protein protects NADH: Ubiquinone oxidoreductase of the electron transport chain during heat stress in plants. FEBS Lett. 1998, 430, 246–250. [Google Scholar] [CrossRef]
  77. Boveris, A.; Oshino, N.; Chance, B. The cellular production of hydrogen peroxide. Biochem. J. 1972, 128, 617–630. [Google Scholar] [CrossRef]
  78. Bu, D.; Bionaz, M.; Wang, M.; Nan, X.; Ma, L.; Wang, J. Transcriptome difference and potential crosstalk between liver and mammary tissue in mid-lactation primiparous dairy cows. PLoS ONE 2017, 12, e0173082. [Google Scholar] [CrossRef]
  79. Baumgard, L.; Wheelock, J.; Sanders, S.; Moore, C.; Green, H.; Waldron, M.; Rhoads, R. Postabsorptive carbohydrate adaptations to heat stress and monensin supplementation in lactating Holstein cows. J. Dairy Sci. 2011, 94, 5620–5633. [Google Scholar] [CrossRef] [PubMed]
  80. Zhang, M.; Zhao, S.; Wang, S.; Luo, C.; Gao, H.; Zheng, N.; Wang, J. d-Glucose and amino acid deficiency inhibits casein synthesis through JAK2/STAT5 and AMPK/mTOR signaling pathways in mammary epithelial cells of dairy cows. J. Dairy Sci. 2018, 101, 1737–1746. [Google Scholar] [CrossRef]
  81. Bu, D.; Ma, L.; Gao, S.T.; Baumgard, L.H.; Bionaz, M. Heat stress decreases transcription of protein metabolism-related genes in mammary tissue of middle lactating cows. J. Dairy Sci. 2017, 100 (Suppl. 2), 223. [Google Scholar]
  82. Elmore, S. Apoptosis: A review of programmed cell death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef] [PubMed]
  83. Ashe, P.C.; Berry, M.D. Apoptotic signaling cascades. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2003, 27, 199–214. [Google Scholar] [CrossRef]
  84. Wang, W.-H.; Grégori, G.; Hullinger, R.L.; Andrisani, O.M. Sustained activation of p38 mitogen-activated protein kinase and c-Jun N-terminal kinase pathways by hepatitis B virus X protein mediates apoptosis via induction of Fas/FasL and tumor necrosis factor (TNF) receptor 1/TNF-α expression. Mol. Cell. Biol. 2004, 24, 10352–10365. [Google Scholar] [CrossRef] [Green Version]
  85. Spierings, D.; McStay, G.; Saleh, M.; Bender, C.; Chipuk, J.; Maurer, U.; Green, D.R. Connected to death: The (unexpurgated) mitochondrial pathway of apoptosis. Science 2005, 310, 66–67. [Google Scholar] [CrossRef] [Green Version]
  86. Belhadj Slimen, I.; Najar, T.; Ghram, A.; Dabbebi, H.; Ben Mrad, M.; Abdrabbah, M. Reactive oxygen species, heat stress and oxidative-induced mitochondrial damage. A review. Int. J. Hyperth. 2014, 30, 513–523. [Google Scholar] [CrossRef]
  87. Rodríguez-Luccioni, H.L.; Latorre-Esteves, M.; Méndez-Vega, J.; Soto, O.; Rodríguez, A.R.; Rinaldi, C.; Torres-Lugo, M. Enhanced reduction in cell viability by hyperthermia induced by magnetic nanoparticles. Int. J. Nanomed. 2011, 6, 373. [Google Scholar]
  88. Reed, J.C. Bcl-2 family proteins. Oncogene 1998, 17, 3225–3236. [Google Scholar] [CrossRef] [Green Version]
  89. Green, D.R.; Kroemer, G. The pathophysiology of mitochondrial cell death. Science 2004, 305, 626–629. [Google Scholar] [CrossRef]
  90. Antonsson, B.; Montessuit, S.; Lauper, S.; Eskes, R.; Martinou, J.-C. Bax oligomerization is required for channel-forming activity in liposomes and to trigger cytochrome c release from mitochondria. Biochem. J. 2000, 345, 271–278. [Google Scholar] [CrossRef] [PubMed]
  91. Wei, M.C.; Zong, W.-X.; Cheng, E.H.-Y.; Lindsten, T.; Panoutsakopoulou, V.; Ross, A.J.; Roth, K.A.; MacGregor, G.R.; Thompson, C.B.; Korsmeyer, S.J. Proapoptotic BAX and BAK: A requisite gateway to mitochondrial dysfunction and death. Science 2001, 292, 727–730. [Google Scholar] [CrossRef] [Green Version]
  92. Shimizu, S.; Matsuoka, Y.; Shinohara, Y.; Yoneda, Y.; Tsujimoto, Y. Essential role of voltage-dependent anion channel in various forms of apoptosis in mammalian cells. J. Cell Biol. 2001, 152, 237–250. [Google Scholar] [CrossRef]
  93. Crompton, M. The mitochondrial permeability transition pore and its role in cell death. Biochem. J. 1999, 341, 233–249. [Google Scholar] [CrossRef] [PubMed]
  94. Petronilli, V.; Penzo, D.; Scorrano, L.; Bernardi, P.; Di Lisa, F. The mitochondrial permeability transition, release of cytochrome c and cell death correlation with the duration of pore openings in situ. J. Biol. Chem. 2001, 276, 12030–12034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Li, J.-J.; Oberley, L.W. Overexpression of manganese-containing superoxide dismutase confers resistance to the cytotoxicity of tumor necrosis factor α and/or hyperthermia. Cancer Res. 1997, 57, 1991–1998. [Google Scholar]
  96. Zhao, Q.-L.; Fujiwara, Y.; Kondo, T. Mechanism of cell death induction by nitroxide and hyperthermia. Free Radic. Biol. Med. 2006, 40, 1131–1143. [Google Scholar] [CrossRef]
  97. Ott, M.; Robertson, J.D.; Gogvadze, V.; Zhivotovsky, B.; Orrenius, S. Cytochrome c release from mitochondria proceeds by a two-step process. Proc. Natl. Acad. Sci. USA 2002, 99, 1259–1263. [Google Scholar] [CrossRef] [Green Version]
  98. Circu, M.L.; Moyer, M.P.; Harrison, L.; Aw, T.Y. Contribution of glutathione status to oxidant-induced mitochondrial DNA damage in colonic epithelial cells. Free Radic. Biol. Med. 2009, 47, 1190–1198. [Google Scholar] [CrossRef] [Green Version]
  99. Rachek, L.I.; Yuzefovych, L.V.; LeDoux, S.P.; Julie, N.L.; Wilson, G.L. Troglitazone, but not rosiglitazone, damages mitochondrial DNA and induces mitochondrial dysfunction and cell death in human hepatocytes. Toxicol. Appl. Pharmacol. 2009, 240, 348–354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Ricci, C.; Pastukh, V.; Leonard, J.; Turrens, J.; Wilson, G.; Schaffer, D.; Schaffer, S.W. Mitochondrial DNA damage triggers mitochondrial-superoxide generation and apoptosis. Am. J. Physiol. Cell Physiol. 2008, 294, C413–C422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Polyak, K.; Xia, Y.; Zweier, J.L.; Kinzler, K.W.; Vogelstein, B. A model for p53-induced apoptosis. Nature 1997, 389, 300–305. [Google Scholar] [CrossRef] [PubMed]
  102. Jacobson, M.D.; Raff, M.C. Programmed cell death and Bcl-2 protection in very low oxygen. Nature 1995, 374, 814–816. [Google Scholar] [CrossRef]
  103. Gilbert, H.F. Biological disulfides: The third messenger? Modulation of phosphofructokinase activity by thiol/disulfide exchange. J. Biol. Chem. 1982, 257, 12086–12091. [Google Scholar] [CrossRef]
  104. Ghibelli, L.; Coppola, S.; Rotilio, G.; Lafavia, E.; Maresca, V.; Ciriolo, M. Non-oxidative loss of glutathione in apoptosis via GSH extrusion. Biochem. Biophys. Res. Commun. 1995, 216, 313–320. [Google Scholar] [CrossRef]
  105. Chen, W.; Liu, Y.; Zhang, L.; Gu, X.; Liu, G.; Shahid, M.; Gao, J.; Ali, T.; Han, B. Nocardia cyriacigeogica from bovine mastitis induced in vitro apoptosis of bovine mammary epithelial cells via activation of mitochondrial-caspase pathway. Front. Cell. Infect. Microbiol. 2017, 7, 194. [Google Scholar] [CrossRef] [PubMed]
  106. Neville, M.C.; McFadden, T.B.; Forsyth, I. Hormonal regulation of mammary differentiation and milk secretion. J. Mammary Gland Biol. Neoplasia 2002, 7, 49–66. [Google Scholar] [CrossRef]
  107. Wang, X.; Proud, C.G. The mTOR pathway in the control of protein synthesis. Physiology 2006, 21, 362–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Bhaskar, P.T.; Hay, N. The two TORCs and AKT. Dev. Cell 2007, 12, 487–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Hay, N.; Sonenberg, N. Upstream and downstream of mTOR. Genes Dev. 2004, 18, 1926–1945. [Google Scholar] [CrossRef] [Green Version]
  110. Alessi, D.R.; Pearce, L.R.; Garcia-Martinez, J.M. New insights into mTOR signaling: mTORC2 and beyond. Sci. Signal. 2009, 2, pe27. [Google Scholar] [CrossRef]
  111. Gingras, A.-C.; Raught, B.; Sonenberg, N. Regulation of translation initiation by FRAP/mTOR. Genes Dev. 2001, 15, 807–826. [Google Scholar] [CrossRef] [Green Version]
  112. Iadevaia, V.; Huo, Y.; Zhang, Z.; Foster, L.J.; Proud, C.G. Roles of the Mammalian Target of Rapamycin, mTOR, in Controlling Ribosome Biogenesis and Protein Synthesis; Portland Press Ltd.: London, UK, 2012. [Google Scholar]
  113. Kimball, S.R.; Jefferson, L.S. New functions for amino acids: Effects on gene transcription and translation. Am. J. Clin. Nutr. 2006, 83, 500S–507S. [Google Scholar] [CrossRef] [Green Version]
  114. CARLBERG, U.; NILSSON, A.; NYGÅRD, O. Functional properties of phosphorylated elongation factor 2. Eur. J. Biochem. 1990, 191, 639–645. [Google Scholar] [CrossRef] [PubMed]
  115. Redpath, N.T.; Foulstone, E.J.; Proud, C.G. Regulation of translation elongation factor-2 by insulin via a rapamycin-sensitive signalling pathway. EMBO J. 1996, 15, 2291–2297. [Google Scholar] [CrossRef] [PubMed]
  116. Yang, X.; Yang, C.; Farberman, A.; Rideout, T.; De Lange, C.; France, J.; Fan, M. The mammalian target of rapamycin-signaling pathway in regulating metabolism and growth. J. Anim. Sci. 2008, 86, E36–E50. [Google Scholar] [CrossRef]
  117. Patursky-Polischuk, I.; Stolovich-Rain, M.; Hausner-Hanochi, M.; Kasir, J.; Cybulski, N.; Avruch, J.; Rüegg, M.A.; Hall, M.N.; Meyuhas, O. The TSC-mTOR pathway mediates translational activation of TOP mRNAs by insulin largely in a raptor-or rictor-independent manner. Mol. Cell. Biol. 2009, 29, 640–649. [Google Scholar] [CrossRef] [Green Version]
  118. Rius, A.; Appuhamy, J.; Cyriac, J.; Kirovski, D.; Becvar, O.; Escobar, J.; McGilliard, M.; Bequette, B.; Akers, R.; Hanigan, M. Regulation of protein synthesis in mammary glands of lactating dairy cows by starch and amino acids. J. Dairy Sci. 2010, 93, 3114–3127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Choi, K.M.; Barash, I.; Rhoads, R.E. Insulin and prolactin synergistically stimulate β-casein messenger ribonucleic acid translation by cytoplasmic polyadenylation. Mol. Endocrinol. 2004, 18, 1670–1686. [Google Scholar] [CrossRef] [Green Version]
  120. Connor, E.; Meyer, M.; Li, R.; Van Amburgh, M.; Boisclair, Y.; Capuco, A. Regulation of gene expression in the bovine mammary gland by ovarian steroids. J. Dairy Sci. 2007, 90, E55–E65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Menzies, K.K.; Lee, H.J.; Lefèvre, C.; Ormandy, C.J.; Macmillan, K.L.; Nicholas, K.R. Insulin, a key regulator of hormone responsive milk protein synthesis during lactogenesis in murine mammary explants. Funct. Integr. Genom. 2010, 10, 87–95. [Google Scholar] [CrossRef] [PubMed]
  122. Kahn, C.R. Insulin resistance, insulin insensitivity, and insulin unresponsiveness: A necessary distinction. Metabolism 1978, 27, 1893–1902. [Google Scholar] [CrossRef]
  123. Kyriakis, J.M.; Avruch, J. Sounding the alarm: Protein kinase cascades activated by stress and inflammation. J. Biol. Chem. 1996, 271, 24313–24316. [Google Scholar] [CrossRef] [Green Version]
  124. Cohen, P. Dissection of protein kinase cascades that mediate cellular response to cytokines and cellular stress. In Advances in Pharmacology; Elsevier: Amsterdam, The Netherlands, 1996; Volume 36, pp. 15–27. [Google Scholar]
  125. Evans, J.L.; Maddux, B.A.; Goldfine, I.D. The molecular basis for oxidative stress-induced insulin resistance. Antioxid. Redox Signal. 2005, 7, 1040–1052. [Google Scholar] [CrossRef]
  126. Qiao, L.-y.; Goldberg, J.L.; Russell, J.C.; Sun, X.J. Identification of enhanced serine kinase activity in insulin resistance. J. Biol. Chem. 1999, 274, 10625–10632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Zick, Y. Insulin resistance: A phosphorylation-based uncoupling of insulin signaling. Trends Cell Biol. 2001, 11, 437–441. [Google Scholar] [CrossRef]
  128. Potashnik, R.; Bloch-Damti, A.; Bashan, N.; Rudich, A. IRS1 degradation and increased serine phosphorylation cannot predict the degree of metabolic insulin resistance induced by oxidative stress. Diabetologia 2003, 46, 639–648. [Google Scholar] [CrossRef]
  129. Brockman, R.P.; Laarveld, B. Hormonal regulation of metabolism in ruminants; a review. Livest. Prod. Sci. 1986, 14, 313–334. [Google Scholar] [CrossRef]
  130. Hayirli, A. The role of exogenous insulin in the complex of hepatic lipidosis and ketosis associated with insulin resistance phenomenon in postpartum dairy cattle. Vet. Res. Commun. 2006, 30, 749–774. [Google Scholar] [CrossRef] [PubMed]
  131. Yadav, B.; Pandey, V.; Yadav, S.; Singh, Y.; Kumar, V.; Sirohi, R. Effect of misting and wallowing cooling systems on milk yield, blood and physiological variables during heat stress in lactating Murrah buffalo. J. Anim. Sci. Technol. 2016, 58, 1–10. [Google Scholar] [CrossRef] [Green Version]
  132. Poljsak, B. Strategies for reducing or preventing the generation of oxidative stress. Oxidative Med. Cell. Longev. 2011, 2011, 194586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. 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. [Google Scholar] [CrossRef] [Green Version]
  134. Renaudeau, D.; Collin, A.; Yahav, S.; De Basilio, V.; Gourdine, J.-L.; Collier, R. Adaptation to hot climate and strategies to alleviate heat stress in livestock production. Animal 2012, 6, 707–728. [Google Scholar] [CrossRef] [Green Version]
  135. Chen, J.M.; Schütz, K.E.; Tucker, C.B. Cooling cows efficiently with water spray: Behavioral, physiological, and production responses to sprinklers at the feed bunk. J. Dairy Sci. 2016, 99, 4607–4618. [Google Scholar] [CrossRef]
  136. Chen, J.M.; Schütz, K.E.; Tucker, C.B. Cooling cows efficiently with sprinklers: Physiological responses to water spray. J. Dairy Sci. 2015, 98, 6925–6938. [Google Scholar] [CrossRef] [PubMed]
  137. Blackshaw, J.K.; Blackshaw, A. Heat stress in cattle and the effect of shade on production and behaviour: A review. Aust. J. Exp. Agric. 1994, 34, 285–295. [Google Scholar] [CrossRef]
  138. Boonsanit, D.; Chanpongsang, S.; Chaiyabutr, N. Effects of supplemental recombinant bovine somatotropin and mist-fan cooling on the renal tubular handling of sodium in different stages of lactation in crossbred Holstein cattle. Res. Vet. Sci. 2012, 93, 417–426. [Google Scholar] [CrossRef] [PubMed]
  139. Sretenović, L.; Aleksić, S.; Petrović, M.P.; Miščević, B. Nutritional factors influencing improvement of milk and meat quality as well as productive and reproductive parameters of cattle. Biotechnol. Anim. Husb. 2007, 23, 217–226. [Google Scholar] [CrossRef]
  140. Surai, P.F.; Kochish, I.I.; Fisinin, V.I.; Juniper, D.T. Revisiting oxidative stress and the use of organic selenium in dairy cow nutrition. Animals 2019, 9, 462. [Google Scholar] [CrossRef] [Green Version]
  141. Sordillo, L. Nutritional strategies to optimize dairy cattle immunity. J. Dairy Sci. 2016, 99, 4967–4982. [Google Scholar] [CrossRef] [PubMed]
  142. Sun, L.; Gao, S.; Wang, K.; Xu, J.; Sanz-Fernandez, M.; Baumgard, L.; Bu, D. Effects of source on bioavailability of selenium, antioxidant status, and performance in lactating dairy cows during oxidative stress-inducing conditions. J. Dairy Sci. 2019, 102, 311–319. [Google Scholar] [CrossRef] [Green Version]
  143. Sun, L.-H.; Huang, J.-Q.; Deng, J.; Lei, X.G. Avian selenogenome: Response to dietary Se and vitamin E deficiency and supplementation. Poult. Sci. 2019, 98, 4247–4254. [Google Scholar] [CrossRef]
  144. Sun, L.; Wang, F.; Wu, Z.; Ma, L.; Baumrucker, C.; Bu, D. Comparison of Selenium Source in Preventing Oxidative Stress in Bovine Mammary Epithelial Cells. Animals 2020, 10, 842. [Google Scholar] [CrossRef]
  145. Singh, V.K.; Beattie, L.A.; Seed, T.M. Vitamin E: Tocopherols and tocotrienols as potential radiation countermeasures. J. Radiat. Res. 2013, 54, 973–988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Packer, L.; Maguire, J.J.; Mehlhorn, R.J.; Serbinova, E.; Kagan, V.E. Mitochondria and microsomal membranes have a free radical reductase activity that prevents chromanoxyl radical accumulation. Biochem. Biophys. Res. Commun. 1989, 159, 229–235. [Google Scholar] [CrossRef]
  147. Packer, L. Interactions among antioxidants in health and disease: Vitamin E and its redox cycle. Proc. Soc. Exp. Biol. Med. 1992, 200, 271–276. [Google Scholar] [CrossRef] [PubMed]
  148. Weaver, J.; Frederikse, H. CRC Handbook of Chemistry and Physics; CRC Press: Boca Raton, FL, USA, 1977; Volume 76, pp. 12–156. [Google Scholar]
  149. Traber, M.G.; Stevens, J.F. Vitamins C and E: Beneficial effects from a mechanistic perspective. Free Radic. Biol. Med. 2011, 51, 1000–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Pappas, A.C.; Zoidis, E.; Chadio, S.E. Maternal selenium and developmental programming. Antioxidants 2019, 8, 145. [Google Scholar] [CrossRef] [Green Version]
  151. Surai, P.F.; Kochish, I.I.; Fisinin, V.I.; Velichko, O.A. Selenium in poultry nutrition: From sodium selenite to organic selenium sources. J. Poult. Sci. 2017, 0170132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Epp, O.; Ladenstein, R.; Wendel, A. The refined structure of the selenoenzyme glutathione peroxidase at 0.2-nm resolution. Eur. J. Biochem. 1983, 133, 51–69. [Google Scholar] [CrossRef] [PubMed]
  153. Holmgren, A.; Lu, J. Thioredoxin and thioredoxin reductase: Current research with special reference to human disease. Biochem. Biophys. Res. Commun. 2010, 396, 120–124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Yang, C.; Chowdhury, M.; Huo, Y.; Gong, J. Phytogenic compounds as alternatives to in-feed antibiotics: Potentials and challenges in application. Pathogens 2015, 4, 137–156. [Google Scholar] [CrossRef] [Green Version]
  155. Miliauskas, G.; Venskutonis, P.; Van Beek, T. Screening of radical scavenging activity of some medicinal and aromatic plant extracts. Food Chem. 2004, 85, 231–237. [Google Scholar] [CrossRef]
  156. Zhong, R.; Zhou, D. Oxidative stress and role of natural plant derived antioxidants in animal reproduction. J. Integr. Agric. 2013, 12, 1826–1838. [Google Scholar] [CrossRef]
  157. Huang, Q.; Liu, X.; Zhao, G.; Hu, T.; Wang, Y. Potential and challenges of tannins as an alternative to in-feed antibiotics for farm animal production. Anim. Nutr. 2018, 4, 137–150. [Google Scholar] [CrossRef] [PubMed]
  158. Maqsood, S.; Benjakul, S. Comparative studies of four different phenolic compounds on in vitro antioxidative activity and the preventive effect on lipid oxidation of fish oil emulsion and fish mince. Food Chem. 2010, 119, 123–132. [Google Scholar] [CrossRef]
  159. Koleckar, V.; Kubikova, K.; Rehakova, Z.; Kuca, K.; Jun, D.; Jahodar, L.; Opletal, L. Condensed and hydrolysable tannins as antioxidants influencing the health. Mini Rev. Med. Chem. 2008, 8, 436–447. [Google Scholar] [CrossRef]
  160. Ho, K.; Huang, J.; Tsai, C.; Lin, T.; Hsu, Y.; Lin, C. Antioxidant Activity of Tannin Components from Vaccinium vitis-idaea L. J. Pharm. Pharmacol. 1999, 51, 1075–1078. [Google Scholar] [CrossRef] [PubMed]
  161. Bors, W.; Michel, C. Antioxidant capacity of flavanols and gallate esters: Pulse radiolysis studies. Free Radic. Biol. Med. 1999, 27, 1413–1426. [Google Scholar] [CrossRef]
  162. Gonçalves, C.; Dinis, T.; Batista, M.T. Antioxidant properties of proanthocyanidins of Uncaria tomentosa bark decoction: A mechanism for anti-inflammatory activity. Phytochemistry 2005, 66, 89–98. [Google Scholar] [CrossRef] [PubMed]
  163. Halliwell, B. Oxidative stress and cancer: Have we moved forward? Biochem. J. 2007, 401, 1–11. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, Z.; Gao, S.; Ouyang, J.; Ma, L.; Bu, D. Impacts of Heat Stress-Induced Oxidative Stress on the Milk Protein Biosynthesis of Dairy Cows. Animals 2021, 11, 726. https://0-doi-org.brum.beds.ac.uk/10.3390/ani11030726

AMA Style

Guo Z, Gao S, Ouyang J, Ma L, Bu D. Impacts of Heat Stress-Induced Oxidative Stress on the Milk Protein Biosynthesis of Dairy Cows. Animals. 2021; 11(3):726. https://0-doi-org.brum.beds.ac.uk/10.3390/ani11030726

Chicago/Turabian Style

Guo, Zitai, Shengtao Gao, Jialiang Ouyang, Lu Ma, and Dengpan Bu. 2021. "Impacts of Heat Stress-Induced Oxidative Stress on the Milk Protein Biosynthesis of Dairy Cows" Animals 11, no. 3: 726. https://0-doi-org.brum.beds.ac.uk/10.3390/ani11030726

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