Next Article in Journal
Seroprevalence of Tick-Borne Infections in Horses from Northern Italy
Previous Article in Journal
Molecular Evidence Reveals the Sympatric Distribution of Cervus nippon yakushimae and Cervus nippon taiouanus on Jeju Island, South Korea
Previous Article in Special Issue
Exposure to Oxy-Tetracycline Changes Gut Bacterial Community Composition in Rainbow Trout: A Preliminary Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nutritional Modulation, Gut, and Omics Crosstalk in Ruminants

1
Key Lab of Agricultural Animal Genetics, Breeding and Reproduction of Ministry of Education, Huazhong Agriculture University, Wuhan 430070, China
2
Animal Production Department, Faculty of Agriculture, Assuit University, Asyut 71515, Egypt
3
College of Veterinary Medicine, Huazhong Agricultural University, Wuhan 430070, China
4
National Center for International Research on Animal Genetics, Breeding and Reproduction (NCIRAGBR), Huazhong Agricultural University, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
Animals 2022, 12(8), 997; https://doi.org/10.3390/ani12080997
Submission received: 10 January 2022 / Revised: 31 March 2022 / Accepted: 5 April 2022 / Published: 12 April 2022
(This article belongs to the Special Issue Application of “Omics” Technologies in Animal Nutrition)

Abstract

:

Simple Summary

Over the last decade, animal nutrition science has been significantly developed, supported by the great advancements in molecular technologies. For scientists, the present "feedomics and nutrigenomics" era continues to evolve and shape how research is designed, performed, and understood. The new omics interpretations have established a new point of view for the nutrition–gene interaction, integrating more comprehensive findings from animal physiology, molecular genetics, and biochemistry. In the ruminant model, this modern approach addresses rumen microbes as a critical intermediate that can deepen the studies of diet–gut interaction with host genomics. The present review discusses nutrigenomics’ and feedomics’ potential contribution to diminishing the knowledge gap about the DNA cellular activities of different nutrients. It also presents how nutritional management can influence the epigenetic pathway, considering the production type, life stage, and species for more sustainable ruminant nutrition strategies.

Abstract

Ruminant nutrition has significantly revolutionized a new and prodigious molecular approach in livestock sciences over the last decade. Wide-spectrum advances in DNA and RNA technologies and analysis have produced a wealth of data that have shifted the research threshold scheme to a more affluent level. Recently, the published literature has pointed out the nutrient roles in different cellular genomic alterations among different ruminant species, besides the interactions with other factors, such as age, type, and breed. Additionally, it has addressed rumen microbes within the gut health and productivity context, which has made interpreting homogenous evidence more complicated. As a more systematic approach, nutrigenomics can identify how genomics interacts with nutrition and other variables linked to animal performance. Such findings should contribute to crystallizing powerful interpretations correlating feeding management with ruminant production and health through genomics. This review will present a road-mapping discussion of promising trends in ruminant nutrigenomics as a reference for phenotype expression through multi-level omics changes.

1. Introduction

Ruminants are distinctive, influential animal species that have become worthy of attention in human food security marathons. The global population is expected to approach 9.15 billion people by the year 2050 [1], and in turn, global food animal production is expected to rise 2.3% annually, which will require rising production proportions [1,2]. Additionally, rapid population growth has made the animal production situation more critical [3,4], besides the presence of severe environmental changes (climate change, global warming, methane, and greenhouse gas emissions) and natural resource limitations (drought and desertification) [5,6,7,8]. Such emerging threats put pressure on the global situation of the animal protein supply due to the feed resource competition with human food production, which disrupts the sustainability of livestock production systems [9].
The classic animal nutrition approach was traditionally dominated by direct studies that examined the feeding practices related to the production phenotypes. However, this approach could not provide enough knowledge about the nutrient dynamics in the GIT, its effect at the tissue level, and, in turn, its reflection in the animal’s productivity. Additionally, it could not explain the mechanism of action of intermediate metabolites in different cellular activities through different tissue types.
Hence, the advancements in molecular biology, molecular nutrition and physiology, high-throughput technologies, and bioinformatics databases have led to the more powerful inclusion of other studies such as epigenetics, metagenomics, metabolomics, transcriptomics, and proteomics. However, these integration trends focused on the diet’s characteristics, its role in altering metabolism, and its effects on the pathways of other metabolites [10,11]. Thus, debates continue about the best strategies for epigenetic interference applications for determining more precise animal requirements that can guide genetic selection programs. Additionally, many questions have been raised about innovative approaches that broaden our interpretations for more efficient feed resource utilization.
“Omics” refers to methodologies that relate to the knowledge about specific identifiable genes in an animal or the microbiome, genes transcribed to mRNA or proteins, and metabolites present within a particular cell, tissue, organ, fluid, or population [12]. Some of the published literature has drawn attention from traditional nutrition studies toward a closer look at feedomics and nutrigenomics. Feedomics is a field of study that looks at how changes in the diet and gut can affect gene expression, and it is also proposed as the “feed-gut-gene scheme.” In comparison, “nutrigenomics” focuses on nutrient molecules’ role in gene expression and the regulatory mechanisms generally at the cellular level [13].
Recently, feedomics and nutrigenomics have made their way to precisely illustrate the nutritional interventions in animal genetics which can open space for genotypic-tailored feeding studies. This revolutionary approach has focused on how feeds talk to genes and how genes respond, addressing a novel holistic approach and redefining the conventional ruminant nutrition–gene pattern in a broad context. Moreover, animal bioscientists have highlighted the host rumen milieu as a critical intermediate player that controls, regulates, or triggers serial changes by the rumen microbiome’s activities [14,15].
This review discusses feeding and nutrition strategies from a molecular genomics point of view. It introduces a larger framework that places the feed–gut context as the first step toward efficient ruminant nutrition for improving animal health and welfare.

2. Molecular Nutrition–Genomic Interferences

It has been reported that genes alone do not necessarily produce phenotypic traits; various environmental aspects can affect the incidence and degree of trait expression. Nutrition is a principal environmental factor; however, it needs profound genomic enlightenment due to the complexity of feeding-related phenotypes such as feed efficiency [16]. Additionally, our knowledge of which nutritional substrates may impact gene expression is limited. Further, the existing literature analyzes the whole scenario from the diet through the rumen to genes, although the final product is not fully discussed and remains inadequate.
Recently, studies have been published that group genomic feedback with ruminant feeding management and feed formulation. As a result, they have helped to determine more precise nutrient requirements for more sustainable strategies in ruminant production systems. Therefore, trials to understand the genetic response to nutrition have been further complicated and have provided an opportunity for novel research studies that can thoroughly explain the intricate relationship between diet and animal tissue genomics [17,18,19,20,21,22,23,24].
DNA microarrays and gene analysis applications could not prove RNA dynamics, whether mRNA synthesis (transcription) or RNA degradation. Therefore, the preference for RNA-based techniques is attributable to DNA’s existence in both active and inactive or dead cells [25]. However, RNA is dynamically distinguished in participating cells, making RNA a more accurate cellular biomarker. Therefore, RNA-based systems are more precise in omics studies, particularly microbial metabolic activity interpretations [26,27,28,29].

3. The Metabolism Messengers for Gene Regulation

In ruminants, researchers’ main challenge is investigating the relationship between metabolism and genes, tracking molecular pathways that primarily depend on an mRNA transcript methodology [28,30,31]. However, the link between mRNA abundance in the tissue and phenotypic or protein changes in tissue’s gene transcription is not simple because the regulatory pathway for protein synthesis is a multi-stage journey [30,31,32]. Previous studies have established the importance of investigating transcripts depending on the output protein’s significance in regulating or controlling specific metabolic processes [23,24,25,26,27,28,29,30,31,32,33,34,35,36]. Similar works also argued the effect of nutrition on proteomic changes and the feasibility of inducing them in ruminants, which are still scant and surpass the application of these studies to rodent models [37].
As we will discuss later, studies in the literature have reported the potential of some dietary components to affect the cellular metabolism and growth transactions differently through the omics context [38,39,40]. However, such explanations are still unsatisfactory because each dietary factor may have a multi-genomic fingerprint distinguishing some metabolic activities linked to gene expression regulators [41,42,43]. Figure 1 presents the feed characteristics and the potential induction of molecular changes.

4. Tracking the Change Cascade across Gastrointestinal Tissues

Recent trends in feedomics have tracked changes at the feed level or the biochemical level such as an intermediate metabolite, mapping the pattern for multiple mRNA alterations [20,33,44,45,46]. Therefore, it was suggested that there are two paths for feed to start the molecular change cascade. Firstly, the GIT changes are induced by the feed’s physical or biochemical action on the rumen and intestinal tissue. Some physical changes such as papillae development affect absorption, post-absorption, and various metabolism functions [47]. For example, it was reported that 47.5 percent of the critical genes in the rumen epithelial tissues of beef steers are involved in metabolic processes [48]. Secondly, passing the baton to the volatile fatty acids (VFAs) results from the microbiota activity, which acts as a metabolic mediator. The VFA action mainly activates or depresses the specialized transcription factors (TFs) by binding to them (e.g., Figure 2).

4.1. Transcription Factors (TFs)

TFs are functional cellular proteins that manage the gene expression process through binding to target gene regulatory regions (silencer or promoter sequences) on the DNA, sparking gene expression series, and controlling the gene transcription rate [39]. Transcription factors are crucial but not the only mediators in the nutrient–gene scene. Recently, reports have shone the spotlight on the nutrient, mediator, and TF complex that is responsible for launching a later phase of gene upregulation [49]. The second wave of gene expression starts after the upregulation of subsequent TF transcription [49]. Previously, it has been reported that transcription factors may harmonically work in networks of transcription factors that respond to dietary factors [50].
There are various types of transcription factors such as peroxisome proliferator-activated receptors (PPARs), liver X receptors (LXRs), and retinoid X receptors (RXRs). The ligand-dependent nuclear receptors (LdNRs), such as PPARs (α, β, and γ), play a central role in the ruminant model [51]. PPARs are known for their vital cellular functions, such as fatty acid catabolism in skeletal muscle [52], regulation of glucose uptake [39], adipogenic actions [53], and fatty acid oxidation [54]. They are mainly activated by fatty acids, regardless of their source—either the diet or an intermediate metabolite as a ruminal fermentation product [34,38].
The LXR family has major regulatory functions for production traits, such as the two known isoforms α and β which are mainly activated by sterols and fatty acids [55]. For example, LXRα showed regulation capacity for SREBF1 (sterol regulatory element-binding transcription factor 1) expression, a crucial transcription factor regulating milk fat synthesis [55,56]. On the contrary, although it is known that retinoids (9-cis-retinoic acid) are the primary activator of RXR, there are limited data on the potential nutrigenomic effects of vitamin A and derivative retinoids such as 9-cis-retinoic acid through RXRα [57,58].

4.2. DNA Methylation

DNA methylation is a critical epigenetic mechanism that affects gene expression for parent-of-origin traits by methyl group addition without any DNA sequence change, affecting DNA activity. This process occurs by an enzyme group, “DNA methyltransferase (DNMT),” composed of five members: DNMT1, DNMT2, DNMT3A, DNMT3B, and DNMT3L. These enzymes’ mode of action includes methyl group addition to the fifth carbon of cytosine (C) in CpG dinucleotides, forming 5-methylcytosine. The methylation rate strictly correlates with the gene expression extent; hypermethylation of the promoter region depresses the expression level, whereas a low degree of methylation or hypomethylation refers to the active gene expression process [59,60]. Adaptation from parents to offspring is a big part of this mechanism, especially adaptation to various environmental conditions such as heat stress [61], a stimulus such as a change in maternal management [62], physiological state [63], mastitis [64], and milk protein synthesis [65].

4.3. Histone Modification

Histone modification is based on physical conformation to chromatin structure reform by adding or removing functional groups from the N-terminal tails of histone proteins such as H2A, H2B, H3, and H4 [66,67,68]. This conserved protein modification also includes lysine methylation, lysine acetylation, serine/threonine phosphorylation, and ubiquitination [69,70,71]. The majority of histone alterations can regulate the developmental style; the modification in the promoter region results in depressing or activating genes corresponding to different environmental stimuli, such as ultraviolet (UV) or other radiation and chemical carcinogens [72].

4.4. Non-Coding RNA (ncRNAs)

The majority of mammalian genomic DNA is transcribed as non-coding RNAs (ncRNAs) [73], which are initially defined as “junk” [74]. The main effects of ncRNAs range from interfering with mRNA stability to regulating mRNA transcription and translation [75,76]. In ruminants, researchers have shown an increased interest in long non-coding RNAs (lncRNA) and microRNAs (miRNAs), which are well-studied types of non-coding RNA.

4.4.1. Long Non-Coding RNAs (LncRNAs)

LncRNAs refer to RNA transcripts greater than 200 base pairs possessing no protein-coding activity. They have been recently appreciated in physiological processes [77]. However, their examination among ruminants is still limited [27,28,78]. Although the precise acts of lncRNAs are not explicit yet, lncRNAs are reported to have a potential regulatory function in the bovine mammary gland through the pathway of lipid metabolism, fatty acid synthesis [78], and calves’ intestinal growth [27].

4.4.2. MicroRNAs (miRNAs)

MicroRNAs (miRNAs) are non-coding RNAs (18–25 nucleotides) that play an essential role in many physiological processes. Moreover, it is estimated that miRNAs form between 1% and 5% of animal genes and are expected to control at least 60% of genes involved in all cellular activities [79]. The interesting role of miRNAs presents through regulating RNA readiness in the posttranscriptional phase and before translation, affecting protein derivation [80,81,82,83,84]. Additionally, miRNAs are well known for their various significant biological functions, such as adipose tissue regulation [85], proliferation and differentiation of gastrointestinal tissue cells [45,82,86,87,88], mammary gland development [26,33,89,90,91,92], and ovary development [85,93]. Therefore, studying the expression and distribution of miRNAs has attracted interest across a wide range of tissues, aiming to interpret diverse cellular mechanisms, particularly from a pathological perspective [94].
Furthermore, the expression and function of specific miRNAs can be modulated by nutrition. For example, in lactating goats, the expression of 30 miRNAs in the mammary gland was modulated through macronutrient deprivation, where 14 miRNAs were upregulated, and 16 miRNAs were downregulated [53].
Thus, the animal gut’s tracking of an inherited genetic change might take different forms depending on the nutrient or the nature of the feed, shaping gene expression, DNA, and histone modification.

5. Nutrition Influence in Tracking of Epigenetics

Among environmental factors, nutrition can induce desirable epigenetic effects [95,96,97,98] for some traits such as fertility [99,100]. However, diet–epigenetic intervention, or the linkage between nutrients and inheritable changes in DNA base pairs, primarily occurs through chemical regulation mechanisms. As depicted in Figure 3, the potential interaction between environmental conditions and animal status can alter the epigenetic style. Furthermore, broad findings have focused on dietary components and various metabolites as signal messengers for cellular activity in reproductive tissues and organs, as well as their significant effect on reproductive efficiency [101,102,103,104,105]. Additionally, fertility–epigenetic studies supported comprehensive nutritional management as an applicable tracking tool for potential reproduction improvements. In addition, it was reported that fatty acids, especially polyunsaturated types, can alter reproductive performance during different life stages, which is also linked to adipose tissue gene expression [106,107,108,109,110,111,112,113,114].
Conversely, some nutrient substrates may have a contradictory action, even though VFAs play multiple roles across several physiological activities as energy sources and significant transcription factor agonists [39]. Nevertheless, they have been shown to inhibit histone deacetylase mechanisms [115]. Epigenetically, some modifications to DNA base pairs do not change the DNA sequence itself but can shape transgenerational transcription phenotypes (e.g., Table 1).

6. Fetal Programming in the Nutrigenomics Context

Maternal nutritional management has a significant impact, especially in late pregnancy when colostrum is secreted: for instance, selenium supplementation and raised IgG levels in the cattle colostrum [125]. Additionally, in pregnant sheep, changing hay-based diets to corn-based diets in the second half of gestation significantly depressed the expression of (H19, MEG8, PEG1, DLK1, and IGF2R) DNMT genes in the fetus muscles. The expression of these genes was found to be associated with embryonic programming and muscle growth [126].
Moreover, pregnant ewes supplemented with protected methionine in late gestation produced lambs heavier than those produced by non-supplemented ewes [127]. Similarly, treating dairy cows with dietary protected methionine in the late gestation upregulated placental genes that participate in neutral AA and glucose transport, accompanied by higher gene and protein expression of mTOR; this change was also associated with increased calf birth weight [128]. Although some studies have reported that maternal nutrition during late gestation could be a way to change the offspring’s miRNA in beef cattle, there are still some questions about the correlations between colostrum’s miRNAs and their effect on offspring [129,130]. Maternal nutrition substantially affects offspring signaling pathways via regulating transplacental transfers [131] or other diverse pathways [118,119,130,131]. Available nutrients mainly pass through a channel in “the placenta” for launching pathways of signaling such as controlling amino acid transport, as is the case for the mammalian target of rapamycin (mTOR) complex, or the peroxisome proliferator-activated receptor γ (PPARγ), which is the leading influencer of lipid pathway regulation.

7. Nutrigenomics during Newborn Animal’s Life

Although the placenta serves as the primary fetus transfer channel for nutrients and other signaling molecules that pass from mother to newborn, colostrum is thought to be the first super source of active proteins, minerals, and vitamins. However, the placenta barely delivers some bioactive molecules such as immunoglobulins, the chief molecules that can hardly be transferred through it [132]. Therefore, colostrum is the sole source of immunoglobulins that play a crucial part in an animal’s lifespan and passive immunity. Previous studies highlighted growth promoters in colostrum such as insulin-like growth factor (IGF-1) and hormones, focusing on colostrum management and its effects on gut development in neonatal animals [133,134]. In the first week after birth, ingesting colostrum could regulate the expression of T and B cell lineage-specific genes in the intestinal mucosa, in addition to miRNAs and microbial colonization, which may control various mucosal immune changes [135]. While delaying the first colostrum administration after birth reduces IgG transfer in calves [136], calves that ingested colostrum had a higher serum content of amino acids (leucine, valine, and glutamate), which are known for their health benefits and immune expression induction, particularly in the colonic mucosal immune system [137,138].
Among the various colostrum components, the higher content of miRNAs becomes an interesting feature that distinguishes this newborn liquid feed from mature milk that can pass through the milk in bovines [139]. Similarly, miRNA is notable for its immune participation effect on B and T cell differentiation, and interleukin production of macrophages [140]. However, colostrum’s miRNAs drew attention as an active biological component and were remarkably nominated as signaling molecules communicating between the mother and her offspring [141]. Many dairy performance fingerprints mentioned that bta-miR-574, which regulates the leptin receptor, controls the development and lactation of mammary tissue in dairy goats [33]. During lactation, the maternal dietary fat content is suggested to be a fundamental controller of miRNAs in colostrum [141]. However, miRNAs related to lipid metabolism may not be associated with changes in energy sources [142].
The above description is not the whole story of nutrigenomics; studies could not fully discover a central role player—a rumen microbe—which contributes through meta-transcriptomic or meta-proteomic factors. Thus, the rumen microbes may be the critical responders for nutritional change and thus launch another wave of serial changes as we are about to discuss.
The neonatal gut microbial community is a strategic partner in calf health and performance. Since microbial colonization starts from the first days of neonatal animal life, it interacts actively with the first diet in early animal life. Therefore, significant rumen development changes can be found in an age-dependent manner [143]. Additionally, through the microbe–host context, reports have proposed that rumen microbiome changes could regulate neonatal gut development [144]. Dietary promotion of a diverse microbial community is mainly favored as an infection-preventive measure in this sensitive stage. It promotes beneficial bacteria colonization in the small intestine, constraining pathogen microbes’ colonization [145,146]. Newborn calves with a lower incidence of diarrhea and higher growth rates tend to have a higher fecal prevalence of Faecalibacterium, a butyrate-producing strain, and major acetate consumers, which intensify the energy content per mole of the ruminal VFAs. Noteworthy, it plays a partial anti-inflammatory role in Faecalibacterium prausnitzii due to the production of metabolites that further depress NF-kB activity and IL-8 production [147].

8. Feedomics and Nutrigenomics Strategies through Premature Diets

Pre-weaned feeding depends on colostrum, milk replacers, or even whole milk, which passes directly to the abomasum due to the esophageal groove’s existence. As a result, newborn ruminants’ reliance on liquid feeds may limit rumen development [18,148,149]. Furthermore, in the rumen epithelium, MCT1 is the major cellular monocarboxylate transporter (such as SCFAs, lactate, pyruvate, and ketone bodies). MCT1 is mainly responsible for transferring energy sources from the ruminal epithelial cells to the bloodstream and maintaining the intracellular pH [148]. Therefore, MCT1 expression in neonatal ruminants may be influenced by liquid feeds [149], intraluminal SCFA concentrations, or a lower pH value [150]. In beef-producing calves, it was found that adopting a strategy of early weaning (at two months of age) and introducing different diets (high dietary starch) resulted in precocious adiposeness activity present in more intramuscular fat deposition, producing higher-graded carcasses. These dynamics in skeletal muscle tissue activated by the dietary change are mainly coordinated by PPARγ and CCAAT enhancer-binding protein alpha (CEBPA) [151].
Various dietary alterations in premature animals can affect their upcoming production patterns through genomic alteration. For example, starter enrichment (especially for protein content) in neonatal Holstein calves elevated PPARA and cell proliferation gene expression (INSR, FOXO1, AKT3) [152]. Additionally, this change was accompanied by upregulated ketogenic genes (HMGCS2, HMGCL, and BDH1) simultaneous to fatty acid oxidation gene (CPT1A, ACADVL) downregulation, mainly suggesting that early dietary enhancements may be a promising route for promoting energy utilization in the ruminal cell, which results in more significant ruminal development [152]. In addition, changes in the early feeding strategy and style of newborn ruminants may influence rumen development and initiate long-term consequences for lifetime productivity [141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156].

9. Feed Efficiency and Gene Expression

Productive, healthy animals require an adequate intake of tallied and well-balanced diets. Caloric density and nutrient availability are among the controllers of metabolism through gene expression by inducing changes in metabolic regulatory signals, mainly since nutrient supply and hormonal status are strictly linked [103,157]. Moreover, low-feed-intake animals are more vulnerable to several immune responses such as inflammation, liver lesions, and bacterial infection [158]. Additionally, efficient animals are the valuable producer’s target because, economically, this means less feed consumption and lower production costs. Moreover, efficient animals showed further environmental benefits such as lower ammonia emissions [159], 28% less methane [160], and 15% less manure [161].
Feed intake and residual feed intake (RFI) have been used as expressions for feed efficiency measures. However, RFI is calculated as the difference between actual feed intake and estimated feed intake on a maintenance and growth requirement basis. Low-RFI animals are considered efficient, whereas high-RFI animals are considered inefficient. Since it is based on energy intake and requirements contrary to the gain: feed ratio, RFI is unrestricted by growth outlines, making RFI a more reliable feed efficiency measure. Researchers highlighted RFI as a precursor to animal energy intake. This opened possibilities to apply genomic selection to this trait, a moderately heritable trait (0.28 to 0.45), to identify genes associated with various physiological pathways [162,163]. Correspondingly, feed efficiency measures were integrated with selecting feed-efficient animals in time-saving, accurate, and cost-efficient styles [158]. Suggested regulatory genes for energy production linked to RFI were also associated with paracellular permeability, which assists various nutrients’ and SCFAs’ transport [88,148]. In beef cattle, low-RFI animals showed higher expression for a group of genes (TPI1, TECR, COX8A, SLC25A39, PKM2, and SUZ12) that play a part in rumen epithelium morphogenesis through facilitating energy production, needed for tissue development [48].
As feed intake changes, the ruminant GIT reacts differently to pH disruption. The ruminal epithelium responds in various ways; as a short-run response, the molecular adaptation includes greater gene expression and proteins participating in VFA transport actions [162,163]. Therefore, the Na+/H+ exchanger’s activity (such as SLC9A1) tends to show elevated expression, which has been nominated as adaptive molecular-physiological feedback for stabilizing pH through the rumen and omasal epithelium [164,165], and higher expression of Na+/H+ exchangers linked to insulin signaling [166]. Then, physical adaptation follows, through expanding absorptive surfaces by the morphological development of the ruminal epithelium, such as hyperplasia and hypertrophy [10,167]. Some of the discussions provided a molecular understanding of the ruminal epithelial absorptive mechanism in feed-efficient animals. VFA uptake synchronized with absorption and upregulating genes in the ruminal epithelium [168,169]. Upregulation of VFA absorption enhances VFA uptake in ruminal epithelial cells, which results in an increased pH level through an elevation in intracellular hydrogen ions to normalize the intracellular pH status [170,171].

10. Genomic Changes through Dietary Management

Feed restriction protocols have frequently been used to examine intake reduction’s effect, its relation to mRNA abundance in GIT tissues, and potential feeding behavior feedback. Previous studies have shown that feed restriction could downregulate specific gene expressions such as α-lactalbumin (LALBA), which is mainly considered responsible for expressing co-enzymes that participate in lactose synthesis, which explains the milk production decline for restricted feed cows [172]. However, it was reported that during short-term feed deprivation, GIT hormones’ (cholecystokinin and glucagon-like peptide 1) concentration decreased due to mRNA abundance depression of these hormones in the duodenum and ileum [37].
Several studies have shown that dietary energy might play an essential role in how different tissues use other nutrients. For example, dietary energy and propionate production could help bovine mammary tissue make more protein [38]. In addition, previous studies have shown that dietary fatty acids can change cellular behavior. For example, they can change the miRNA regulation of ovine adipogenic genes [173] or make a specific gene more active, which might be an inflammatory mediator such as L-selectin [174]. Furthermore, controlled energy intake also confers ruminant advantages by triggering hepatic molecular adaptations well ahead of parturition [175]. In this connection, intensifying the dietary caloric content using unsaturated fats is more favorable than using oils. This preference for saturated over unsaturated fats in ruminant diets is due to the higher digestibility of saturated than unsaturated forms, which also depress milk fat [176]. It is noteworthy that the abundance of mRNA transcripts in pregnant, repeat-breeding cows that were fed n-3 PUFA-rich diets showed upregulated interferon-stimulated gene (ISG) expression, accompanied by an increased preovulatory follicle (POF) size [21]. Additionally, n-3 PUFA supplementation was correlated with suppressing the pulsatile endometrium secretion of PGF2α that had anti-luteolytic activity [109], besides higher embryonic survival [177].
In the same vein, energy overfeeding of dairy cattle in the dry period has been linked to transcriptional changes, disposing cows to fatty liver, and perhaps overall liver health during the periparturient period [175]. Moreover, it has been conclusively shown that higher-feed-intake beef steers showed significant increases in gene expression responsible for cell growth and proliferation, highlighting factors associated with glycolysis and oxidative phosphorylation in rumen epithelial cells [48].
It is thought that cutting back on food could affect reproductive traits and growth in the small ruminant model. In addition to gut morphology impairments, early feed-restricted ewe lambs showed inefficient reproduction performance and retarded live body weights [178]. Contrarily, this suggestion raises the negative consequence of the acidic effects of high-energy diets that depress cell barrier capacity against various damaging molecules. In addition, an energy-rich diet could weaken some rumen epithelial cellular immunity functions by depressing the expression of some proteins such as HSP71 [18].
Furthermore, various proteins’ abundance and shifts in the genes expressed in the ruminal epithelium showed a linkage to metabolite flux. This abundance, which may be related to changes in ruminal bacterial species [179], was closely related to the metabolite profile, with significantly higher ruminal SCFA concentrations, particularly valerate and butyrate [180,181,182]. Moreover, butyrate, the key influencer in epithelial barrier integration [183], also affects the expression of genes contributing to other SCFA transports through the ruminal epithelium (SLC16A3, SLC26A3, and HIF1A) in sheep and (PAT1, AE2, MCT1, and NHE2) in goats [184,185]. Noteworthy, short-chain fatty acids (propionate specifically) could decelerate GH expression and prolactin (PRL) in dairy cow anterior pituitary cells [186].

11. The Future of Research in GIT Mucosal Immunity

The previous sections integrated the diet–GIT development of neonates with mature animals, linking it with transcriptomic changes. These things are essential to understanding how animal performance can be affected by changes in nutrition through the gut.
The previous results highlighted that the mucosal epithelial architecture change could result in antigen changes and various innate immune responses followed by a disturbance in cytokine profiles. Therefore, diet–microbiota and host immune modulation interventions against gastrointestinal pathogens can significantly optimize production performance and minimize gastrointestinal disease [187]. Additionally, such interventions are mainly responsible for early life stage stress [188], which significantly depresses the newborn animals’ growth performance and health. However, there is a knowledge gap about the mechanisms involved, especially from the host side of this host–diet interaction. Additionally, novel molecular approaches such as fecal microbiome RNA can enrich this research spot and introduce a more deep interpretation of the dynamics of the cellular changes during the different animal life stages [189].
Therefore, the mucosal immune functions have opened future questions about whether the GIT mucosal immunity can be a starting point for re-evaluating nutritional management and strategies, especially for ruminants.

12. Conclusions

Feedomics and nutrigenomics have revolutionized our previous knowledge about ruminant nutrition. The interaction between feeding and gene expression can be manipulated for more benefits concerning animal health, production sustainability, and welfare. DNA- and RNA-based technologies empower researchers to form a comprehensive picture of the feed effect on biological changes, and metabolic and epigenetic mechanisms. Additionally, feedomics and nutrigenomics studies revealed the critical role of the rumen microbiome that is present mightily in many physiological-metabolic pathways. Additional factors must be considered through feedomics studies, such as age, animal species, production phase, and gut–host relations. Different dietary diet/gene connections between production systems are complicated, especially with multi-gene expression changes. In all the studies reviewed here, nutrigenomics insights support researchers in remodeling feeding practices efficiently and isolating diet-induced changes from other causes of change such as age and development. It is hoped that this review will help to build a bigger picture that can show how each dietary component has a unique genomic response that can be used in future feeding management strategies.

Author Contributions

Conceptualization, M.A. and L.Y.; investigation, A.S., M.F.-e.-A.K., H.L., W.W., A.A., Z.Y., M.J.A. and A.L.; writing—original draft preparation, M.A.; writing—review and editing, M.A., A.L. and L.Y.; supervision, L.Y.; project administration, L.Y.; funding acquisition, L.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the China Agriculture Research System of MOF and MARA.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. McLeod, A. Others World Livestock 2011-Livestock in Food Security; Food and Agriculture Organization of the United Nations (FAO): Rome, Italy, 2011. [Google Scholar]
  2. Bouwman, L.; Goldewijk, K.K.; Van Der Hoek, K.W.; Beusen, A.H.W.; Van Vuuren, D.P.; Willems, J.; Rufino, M.C.; Stehfest, E. Exploring global changes in nitrogen and phosphorus cycles in agriculture induced by livestock production over the 1900–2050 period. Proc. Natl. Acad. Sci. USA 2013, 110, 20882–20887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Dangal, S.R.S.; Tian, H.; Zhang, B.; Pan, S.; Lu, C.; Yang, J. Methane emission from global livestock sector during 1890—2014: Magnitude, trends and spatiotemporal patterns. Glob. Chang. Biol. 2017, 23, 4147–4161. [Google Scholar] [CrossRef] [PubMed]
  4. Godber, O.F.; Wall, R. Livestock and food security: Vulnerability to population growth and climate change. Glob. Chang. Biol. 2014, 20, 3092–3102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Svinurai, W.; Mapanda, F.; Sithole, D.; Moyo, E.N.; Ndidzano, K.; Tsiga, A.; Zhakata, W. Enteric methane emissions and their response to agro-ecological and livestock production systems dynamics in Zimbabwe. Sci. Total Environ. 2018, 616–617, 710–719. [Google Scholar] [CrossRef]
  6. Mekonnen, M.M.; Neale, C.M.U.; Ray, C.; Erickson, G.E.; Hoekstra, A.Y. Water productivity in meat and milk production in the US from 1960 to 2016. Environ. Int. 2019, 132, 105084. [Google Scholar] [CrossRef]
  7. Teague, W.R. Forages and pastures symposium: Cover crops in livestock production: Whole-system approach: Managing grazing to restore soil health and farm livelihoods. J. Anim. Sci. 2018, 96, 1519–1530. [Google Scholar] [CrossRef] [Green Version]
  8. Soteriades, A.D.; Foskolos, A.; Styles, D.; Gibbons, J.M. Diversification not specialization reduces global and local environmental burdens from livestock production. Environ. Int. 2019, 132, 104837. [Google Scholar] [CrossRef]
  9. Salter, A.M. Improving the sustainability of global meat and milk production. Proc. Nutr. Soc. 2017, 76, 22–27. [Google Scholar] [CrossRef] [Green Version]
  10. Sun, H.Z.; Guan, L.L. Feedomics: Promises for food security with sustainable food animal production. TrAC Trends Anal. Chem. 2018, 107, 130–141. [Google Scholar] [CrossRef]
  11. Sun, H.Z.; Plastow, G.; Guan, L.L. Invited review: Advances and challenges in application of feedomics to improve dairy cow production and health. J. Dairy Sci. 2019, 102, 5853–5870. [Google Scholar] [CrossRef]
  12. Durkin, L.A.; Childs, C.E.; Calder, P.C. Omega-3 polyunsaturated fatty acids and the intestinal epithelium—A review. Foods 2021, 10, 199. [Google Scholar] [CrossRef] [PubMed]
  13. Loor, J.J.; Bionaz, M.; Drackley, J.K. Systems physiology in dairy cattle: Nutritional genomics and beyond. Annu. Rev. Anim. Biosci. 2013, 1, 365–392. [Google Scholar] [CrossRef]
  14. Xue, M.-Y.; Sun, H.-Z.; Wu, X.-H.; Liu, J.-X.; Guan, L.L. Multi-omics reveals that the rumen microbiome and its metabolome together with the host metabolome contribute to individualized dairy cow performance. Microbiome 2020, 8, 64. [Google Scholar] [CrossRef] [PubMed]
  15. O’Hara, E.; Neves, A.L.A.; Song, Y.; Guan, L.L. The Role of the Gut Microbiome in Cattle Production and Health: Driver or Passenger? Annu. Rev. Anim. Biosci. 2020, 8, 199–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Li, S.; Wang, Q.; Lin, X.; Jin, X.; Liu, L.; Wang, C.; Chen, Q.; Liu, J.; Liu, H. The Use of “Omics” in Lactation Research in Dairy Cows. Int. J. Mol. Sci. 2017, 18, 983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Tamate, H.; McGilliard, A.D.; Jacobson, N.L.; Getty, R. Effect of Various Dietaries on the Anatomical Development of the Stomach in the Calf. J. Dairy Sci. 1962, 45, 408–420. [Google Scholar] [CrossRef]
  18. Hollmann, M.; Miller, I.; Hummel, K.; Sabitzer, S.; Metzler-Zebeli, B.U.; Razzazi-Fazeli, E.; Zebeli, Q. Downregulation of cellular protective factors of rumen epithelium in goats fed high energy diet. PLoS ONE 2013, 8, e81602. [Google Scholar] [CrossRef]
  19. Zhang, H.; Ao, C.J.; Song, L.W.; Zhang, X.F. Effects of different model diets on milk composition and expression of genes related to fatty acid synthesis in the mammary gland of lactating dairy goats. J. Dairy Sci. 2015, 98, 4619–4628. [Google Scholar] [CrossRef] [Green Version]
  20. Zhang, Y.; Yang, H.; Han, L.; Li, F.; Zhang, T.; Pang, J.; Feng, X.; Ren, C.; Mao, S.; Wang, F. Long noncoding RNA expression profile changes associated with dietary energy in the sheep testis during sexual maturation. Sci. Rep. 2017, 7, 5180. [Google Scholar] [CrossRef] [Green Version]
  21. Teeli, A.S.; Sheikh, P.A.; Patra, M.K.; Singh, D.; Kumar, B.; Kumar, H.; Singh, S.K.; Verma, M.R.; Krishnaswamy, N. Effect of dietary n-3 polyunsaturated rich fish oil supplementation on ovarian function and interferon stimulated genes in the repeat breeding cow. Anim. Reprod. Sci. 2019, 211, 106230. [Google Scholar] [CrossRef]
  22. Wilkens, M.R.; Firmenich, C.S.; Schnepel, N.; Muscher-Banse, A.S. A reduced protein diet modulates enzymes of vitamin D and cholesterol metabolism in young ruminants. J. Steroid Biochem. Mol. Biol. 2019, 186, 196–202. [Google Scholar] [CrossRef] [PubMed]
  23. Yohe, T.T.; Schramm, H.; White, R.R.; Hanigan, M.D.; Parsons, C.L.M.; Tucker, H.L.M.; Enger, B.D.; Hardy, N.R.; Daniels, K.M. Form of calf diet and the rumen. II: Impact on volatile fatty acid absorption. J. Dairy Sci. 2019, 102, 8502–8512. [Google Scholar] [CrossRef] [PubMed]
  24. Wilkens, M.R.; Schnepel, N.; Muscher-Banse, A.S. Dietary protein and calcium modulate parathyroid vitamin D receptor expression in young ruminants. J. Steroid Biochem. Mol. Biol. 2020, 196, 105503. [Google Scholar] [CrossRef] [PubMed]
  25. Lettat, A.; Benchaar, C. Diet-induced alterations in total and metabolically active microbes within the rumen of dairy cows. PLoS ONE 2013, 8, e60978. [Google Scholar]
  26. Ibeagha-Awemu, E.M.; Li, R.; Dudemaine, P.-L.; Do, D.N.; Bissonnette, N. Transcriptome analysis of long non-coding RNA in the bovine mammary gland following dietary supplementation with linseed oil and safflower oil. Int. J. Mol. Sci. 2018, 19, 3610. [Google Scholar] [CrossRef] [Green Version]
  27. Ibeagha-Awemu, E.M.; Do, D.N.; Dudemaine, P.-L.; Fomenky, B.E.; Bissonnette, N. Integration of lncRNA and mRNA transcriptome analyses reveals genes and pathways potentially involved in calf intestinal growth and development during the early weeks of life. Genes 2018, 9, 142. [Google Scholar] [CrossRef] [Green Version]
  28. Arora, R.; Sharma, A.; Sharma, U.; Girdhar, Y.; Kaur, M.; Kapoor, P.; Ahlawat, S.; Vijh, R.K. Buffalo milk transcriptome: A comparative analysis of early, mid and late lactation. Sci. Rep. 2019, 9, 5993. [Google Scholar] [CrossRef] [Green Version]
  29. Yang, B.; Chen, H.; Cao, J.; He, B.; Wang, S.; Luo, Y.; Wang, J. Transcriptome Analysis Reveals That Alfalfa Promotes Rumen Development through Enhanced Metabolic Processes and Calcium Transduction in Hu Lambs. Front. Genet. 2019, 10, 929. [Google Scholar] [CrossRef] [Green Version]
  30. E Hernandez-Castellano, L.; M Almeida, A.; Castro, N.; Arguello, A. The colostrum proteome, ruminant nutrition and immunity: A review. Curr. Protein Pept. Sci. 2014, 15, 64–74. [Google Scholar] [CrossRef]
  31. Wu, X.; Sun, H.; Xue, M.; Wang, D.; Guan, L.L.; Liu, J. Serum metabolome profiling revealed potential biomarkers for milk protein yield in dairy cows. J. Proteom. 2018, 184, 54–61. [Google Scholar] [CrossRef]
  32. Shahzad, K.; J. Loor, J. Application of Top-Down and Bottom-up Systems Approaches in Ruminant Physiology and Metabolism. Curr. Genom. 2012, 13, 379–394. [Google Scholar] [CrossRef] [Green Version]
  33. Hou, J.; An, X.; Song, Y.; Cao, B.; Yang, H.; Zhang, Z.; Shen, W.; Li, Y. Detection and comparison of microRNAs in the caprine mammary gland tissues of colostrum and common milk stages. BMC Genet. 2017, 18, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Osorio, J.S.; Vailati-Riboni, M.; Palladino, A.; Luo, J.; Loor, J.J. Application of nutrigenomics in small ruminants: Lactation, growth, and beyond. Small Rumin. Res. 2017, 154, 29–44. [Google Scholar] [CrossRef]
  35. Krishnan, B.B.; Selvaraju, S.; Gowda, N.K.S.; Subramanya, K.B.; Pal, D.; Archana, S.S.; Bhatta, R. Dietary boron supplementation enhances sperm quality and immunity through influencing the associated biochemical parameters and modulating the genes expression at testicular tissue. J. Trace Elem. Med. Biol. 2019, 55, 6–14. [Google Scholar] [CrossRef]
  36. Restelli, L.; Marques, A.T.; Savoini, G.; Invernizzi, G.; Carisetti, M.; Lecchi, C.; Bendixen, E.; Ceciliani, F. Saturated or unsaturated fat supplemented maternal diets influence omental adipose tissue proteome of suckling goat-kids. Res. Vet. Sci. 2019, 125, 451–458. [Google Scholar] [CrossRef] [PubMed]
  37. Drackley, J.K.; Donkin, S.S.; Reynolds, C.K. Major advances in fundamental dairy cattle nutrition. J. Dairy Sci. 2006, 89, 1324–1336. [Google Scholar] [CrossRef]
  38. Bionaz, M.; Chen, S.; Khan, M.J.; Loor, J.J. Functional role of PPARs in ruminants: Potential targets for fine-tuning metabolism during growth and lactation. PPAR Res. 2013, 2013, 684159. [Google Scholar] [CrossRef] [Green Version]
  39. Osorio, J.S.; Moisa, S.J. Gene Regulation in Ruminants: A Nutritional Perspective. In Gene Expression and Control; IntechOpen: London, UK, 2019. [Google Scholar]
  40. Pitta, D.W.; Indugu, N.; Baker, L.; Vecchiarelli, B.; Attwood, G. Symposium review: Understanding diet–microbe interactions to enhance productivity of dairy cows. J. Dairy Sci. 2018, 101, 7661–7679. [Google Scholar] [CrossRef] [Green Version]
  41. Bauman, D.E.; Harvatine, K.J.; Lock, A.L. Nutrigenomics, rumen-derived bioactive fatty acids, and the regulation of milk fat synthesis. Annu. Rev. Nutr. 2011, 31, 299–319. [Google Scholar] [CrossRef]
  42. Tremblay, B.L.; Rudkowska, I. Nutrigenomic point of view on effects and mechanisms of action of ruminant trans fatty acids on insulin resistance and type 2 diabetes. Nutr. Rev. 2017, 75, 214–223. [Google Scholar] [CrossRef]
  43. Nayeri, S.; Stothard, P. Tissues, Metabolic Pathways and Genes of Key Importance in Lactating Dairy Cattle. Springer Sci. Rev. 2016, 4, 49–77. [Google Scholar] [CrossRef] [Green Version]
  44. Green, C.D.; Huang, Y.; Dou, X.; Yang, L.; Liu, Y.; Han, J.D.J. Impact of Dietary Interventions on Noncoding RNA Networks and mRNAs Encoding Chromatin-Related Factors. Cell Rep. 2017, 18, 2957–2968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Steele, M.A.; Penner, G.B.; Chaucheyras-Durand, F.; Guan, L.L. Development and physiology of the rumen and the lower gut: Targets for improving gut health. J. Dairy Sci. 2016, 99, 4955–4966. [Google Scholar] [CrossRef] [Green Version]
  46. Caroprese, M.; Giannenas, I.; Fthenakis, G.C. Interactions between nutritional approaches and defences against microbial diseases in small ruminants. Vet. Microbiol. 2015, 181, 8–14. [Google Scholar] [CrossRef] [PubMed]
  47. Novak, T.E.; Rodriguez-Zas, S.L.; Southey, B.R.; Starkey, J.D.; Stockler, R.M.; Alfaro, G.F.; Moisá, S.J. Jersey steer ruminal papillae histology and nutrigenomics with diet changes. J. Anim. Physiol. Anim. Nutr. 2019, 103, 1694–1707. [Google Scholar] [CrossRef]
  48. Kong, R.S.G.; Liang, G.; Chen, Y.; Stothard, P. Transcriptome profiling of the rumen epithelium of beef cattle differing in residual feed intake. BMC Genom. 2016, 17, 592. [Google Scholar] [CrossRef] [Green Version]
  49. Rakhshandehroo, M.; Knoch, B.; Müller, M.; Kersten, S. Peroxisome proliferator-activated receptor alpha target genes. PPAR Res. 2010, 2010, 612089. [Google Scholar] [CrossRef] [Green Version]
  50. Osorio, J.S.; Lohakare, J.; Bionaz, M. Biosynthesis of milk fat, protein, and lactose: Roles of transcriptional and posttranscriptional regulation. Physiol. Genom. 2016, 48, 231–256. [Google Scholar] [CrossRef] [Green Version]
  51. Burris, T.P.; Solt, L.A.; Wang, Y.; Crumbley, C.; Banerjee, S.; Griffett, K.; Lundasen, T.; Hughes, T.; Kojetin, D.J. Nuclear Receptors and Their Selective Pharmacologic Modulators. Pharmacol. Rev. 2013, 65, 710–778. [Google Scholar] [CrossRef] [Green Version]
  52. Desvergne, B.; Michalik, L.; Wahli, W. Transcriptional regulation of metabolism. Physiol. Rev. 2006, 86, 465–514. [Google Scholar] [CrossRef]
  53. Moisá, S.J.; Shike, D.W.; Meteer, W.T.; Keisler, D.; Faulkner, D.B.; Loor, J.J. Yin yang 1 and adipogenic gene network expression in longissimus muscle of beef cattle in response to nutritional management. Gene Regul. Syst. Bio. 2013, 7, 71–83. [Google Scholar] [CrossRef] [PubMed]
  54. Agazzi, A.; Invernizzi, G.; Campagnoli, A.; Ferroni, M.; Fanelli, A.; Cattaneo, D.; Galmozzi, A.; Crestani, M.; Dell’Orto, V.; Savoini, G. Effect of different dietary fats on hepatic gene expression in transition dairy goats. Small Rumin. Res. 2010, 93, 31–40. [Google Scholar] [CrossRef]
  55. Wang, W.; Luo, J.; Zhong, Y.; Lin, X.-Z.; Shi, H.-B.; Zhu, J.-J.; Li, J.; Sun, Y.-T.; Zhao, W.-S. Goat liver X receptor α, molecular cloning, functional characterization and regulating fatty acid synthesis in epithelial cells of goat mammary glands. Gene 2012, 505, 114–120. [Google Scholar] [CrossRef] [PubMed]
  56. McFadden, J.W.; Corl, B.A. Activation of liver X receptor (LXR) enhances de novo fatty acid synthesis in bovine mammary epithelial cells. J. Dairy Sci. 2010, 93, 4651–4658. [Google Scholar] [CrossRef] [PubMed]
  57. Akbar, H.; Schmitt, E.; Ballou, M.A.; Corrêa, M.N.; DePeters, E.J.; Loor, J.J. Dietary lipid during late-pregnancy and early-lactation to manipulate metabolic and inflammatory gene network expression in dairy cattle liver with a focus on PPARs. Gene Regul. Syst. Bio. 2013, 7, 103–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Goszczynski, D.E.; Mazzucco, J.P.; Ripoli, M.V.; Villarreal, E.L.; Rogberg-Muñoz, A.; Mezzadra, C.A.; Melucci, L.M.; Giovambattista, G. Genetic characterisation of PPARG, CEBPA and RXRA, and their influence on meat quality traits in cattle. J. Anim. Sci. Technol. 2016, 58, 14. [Google Scholar] [CrossRef] [Green Version]
  59. Bionaz, M.; Osorio, J.; Loor, J.J. TRIENNIAL LACTATION SYMPOSIUM: Nutrigenomics in dairy cows: Nutrients, transcription factors, and techniques. J. Anim. Sci. 2015, 93, 5531–5553. [Google Scholar] [CrossRef] [Green Version]
  60. González-Recio, O.; Ugarte, E.; Bach, A. Trans-Generational Effect of Maternal Lactation during Pregnancy: A Holstein Cow Model. PLoS ONE 2012, 7, e51816. [Google Scholar] [CrossRef]
  61. De Barros, F.R.O.; Paula-Lopes, F.F. Cellular and epigenetic changes induced by heat stress in bovine preimplantation embryos. Mol. Reprod. Dev. 2018, 85, 810–820. [Google Scholar] [CrossRef] [Green Version]
  62. Engmann, O. Dairy cows—An opportunity in the research field of non-genetic inheritance? Environ. Epigenetics 2018, 4, dvy014. [Google Scholar] [CrossRef]
  63. Nguyen, M.; Boutinaud, M.; Pétridou, B.; Gabory, A.; Pannetier, M.; Chat, S.; Bouet, S.; Jouneau, L.; Jaffrezic, F.; Laloë, D. DNA methylation and transcription in a distal region upstream from the bovine AlphaS1 casein gene after once or twice daily milking. PLoS ONE 2014, 9, e111556. [Google Scholar] [CrossRef] [PubMed]
  64. Vanselow, J.; Yang, W.; Herrmann, J.; Zerbe, H.; Schuberth, H.-J.; Petzl, W.; Tomek, W.; Seyfert, H.-M. DNA-remethylation around a STAT5-binding enhancer in the αS1-casein promoter is associated with abrupt shutdown of αS1-casein synthesis during acute mastitis. J. Mol. Endocrinol. 2006, 37, 463–477. [Google Scholar] [CrossRef] [PubMed]
  65. Tian, P.; Luo, Y.; Li, X.; Tian, J.; Tao, S.; Hua, C.; Geng, Y.; Ni, Y.; Zhao, R. Negative effects of long-term feeding of high-grain diets to lactating goats on milk fat production and composition by regulating gene expression and DNA methylation in the mammary gland. J. Anim. Sci. Biotechnol. 2017, 8, 74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Hake, S.B.; Xiao, A.; Allis, C.D. Linking the epigenetic ‘language’ of covalent histone modifications to cancer. Br. J. Cancer 2004, 90, 761–769. [Google Scholar] [CrossRef] [Green Version]
  67. Kouzarides, T. Chromatin Modifications and Their Function. Cell 2007, 128, 693–705. [Google Scholar] [CrossRef] [Green Version]
  68. Canani, R.B.; Di Costanzo, M.; Leone, L.; Bedogni, G.; Brambilla, P.; Cianfarani, S.; Nobili, V.; Pietrobelli, A.; Agostoni, C. Epigenetic mechanisms elicited by nutrition in early life. Nutr. Res. Rev. 2011, 24, 198–205. [Google Scholar] [CrossRef]
  69. Tan, M.; Luo, H.; Lee, S.; Jin, F.; Yang, J.S.; Montellier, E.; Buchou, T.; Cheng, Z.; Rousseaux, S.; Rajagopal, N.; et al. Identification of 67 Histone Marks and Histone Lysine Crotonylation as a New Type of Histone Modification. Cell 2011, 146, 1016–1028. [Google Scholar] [CrossRef] [Green Version]
  70. Sadakierska-Chudy, A.; Filip, M. A comprehensive view of the epigenetic landscape. Part II: Histone post-translational modification, nucleosome level, and chromatin regulation by ncRNAs. Neurotox. Res. 2015, 27, 172–197. [Google Scholar] [CrossRef] [Green Version]
  71. Brockers, K.; Schneider, R. Histone H1, the forgotten histone. Epigenomics 2019, 11, 363–366. [Google Scholar] [CrossRef] [Green Version]
  72. Oberdoerffer, P.; Sinclair, D.A. The role of nuclear architecture in genomic instability and ageing. Nat. Rev. Mol. Cell Biol. 2007, 8, 692–702. [Google Scholar] [CrossRef]
  73. Mercer, T.R.; Dinger, M.E.; Sunkin, S.M.; Mehler, M.F.; Mattick, J.S. Specific expression of long noncoding RNAs in the mouse brain. Proc. Natl. Acad. Sci. USA 2008, 105, 716–721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Loomis, W.F.; Gilpin, M.E. Multigene families and vestigial sequences. Proc. Natl. Acad. Sci. USA 1986, 83, 2143–2147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Mercer, T.R.; Mattick, J.S. Structure and function of long noncoding RNAs in epigenetic regulation. Nat. Struct. Mol. Biol. 2013, 20, 300–307. [Google Scholar] [CrossRef]
  76. Heard, E.; Martienssen, R.A. Transgenerational Epigenetic Inheritance: Myths and Mechanisms. Cell 2014, 157, 95–109. [Google Scholar] [CrossRef] [Green Version]
  77. Vance, K.W.; Ponting, C.P. Transcriptional regulatory functions of nuclear long noncoding RNAs. Trends Genet. 2014, 30, 348–355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Ibeagha-Awemu, E.M.; Li, R.; Dudemaine, P.L. The long non-coding RNA transcriptome of the bovine mammary gland and potential regulatory roles in fatty acid synthesis. In Proceedings of the 6th International Symposium on Animal Functional Genomics (6th ISFAG), Piacenza, Italy, 27–29 July 2015; Volume 91. [Google Scholar]
  79. Sohel, M.M.H. Macronutrient modulation of mRNA and microRNA function in animals: A review. Anim. Nutr. 2020, 6, 258–268. [Google Scholar] [CrossRef]
  80. Lim, L.P.; Lau, N.C.; Garrett-Engele, P.; Grimson, A.; Schelter, J.M.; Castle, J.; Bartel, D.P.; Linsley, P.S.; Johnson, J.M. Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature 2005, 433, 769–773. [Google Scholar] [CrossRef]
  81. García-Segura, L.; Pérez-Andrade, M.; Miranda-Ríos, J. The emerging role of MicroRNAs in the regulation of gene expression by nutrients. Lifestyle Genom. 2013, 6, 16–31. [Google Scholar] [CrossRef]
  82. Liang, G.; Malmuthuge, N.; Griebel, P. Model systems to analyze the role of miRNAs and commensal microflora in bovine mucosal immune system development. Mol. Immunol. 2015, 66, 57–67. [Google Scholar] [CrossRef]
  83. Liao, R.; Lv, Y.; Zhu, L.; Lin, Y. Altered expression of miRNAs and mRNAs reveals the potential regulatory role of miRNAs in the developmental process of early weaned goats. PLoS ONE 2019, 14, e0220907. [Google Scholar] [CrossRef] [Green Version]
  84. Quan, S.; Xiong, B.; Nan, X.; Wang, K.; Jiang, L.; Junhu, Y. Different diets change the expression of bovine serum extracellular vesicle-miRNAs. Animals 2019, 9, 1137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Romao, J.M.; Jin, W.; He, M.; McAllister, T. MicroRNAs in bovine adipogenesis: Genomic context, expression and function. BMC Genom. 2014, 15, 137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Aluwong, T.; Kobo, P.I.; Abdullahi, A. Volatile fatty acids production in ruminants and the role of monocarboxylate transporters: A review. African J. Biotechnol. 2010, 9, 6229–6232. [Google Scholar] [CrossRef]
  87. Liang, G.; Malmuthuge, N.; McFadden, T.B.; Bao, H.; Griebel, P.J.; Stothard, P. Potential regulatory role of microRNAs in the development of bovine gastrointestinal tract during early life. PLoS ONE 2014, 9, e92592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Shen, H.; Xu, Z.; Shen, Z.; Lu, Z. The regulation of ruminal short-chain fatty acids on the functions of rumen barriers. Front. Physiol. 2019, 10, 1305. [Google Scholar] [CrossRef]
  89. Ferreira, A.M.; Bislev, S.L.; Bendixen, E.; Almeida, A.M. ScienceDirect The mammary gland in domestic ruminants: A systems biology perspective. J. Proteom. 2013, 94, 110–123. [Google Scholar] [CrossRef] [PubMed]
  90. Weller, M.; Albino, R.L.; Marcondes, M.I.; Silva, W.; Daniels, K.M.; Campos, M.M.; Duarte, M.S.; Mescouto, M.L.; Silva, F.F.; Guimarães, S.E.F. Effects of nutrient intake level on mammary parenchyma growth and gene expression in crossbred (Holstein× Gyr) prepubertal heifers. J. Dairy Sci. 2016, 99, 9962–9973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Tang, K.Q.; Wang, Y.N.; Zan, L.S.; Yang, W.C. miR-27a controls triacylglycerol synthesis in bovine mammary epithelial cells by targeting peroxisome proliferator-activated receptor gamma. J. Dairy Sci. 2017, 100, 4102–4112. [Google Scholar] [CrossRef]
  92. Hare, K.S.; Leal, L.N.; Romao, J.M.; Hooiveld, G.J.; Soberon, F.; Berends, H.; Van Amburgh, M.E.; Martín-Tereso, J.; Steele, M.A. Preweaning nutrient supply alters mammary gland transcriptome expression relating to morphology, lipid accumulation, DNA synthesis, and RNA expression in Holstein heifer calves. J. Dairy Sci. 2019, 102, 2618–2630. [Google Scholar] [CrossRef]
  93. Singh, P.; Golla, N.; Singh, P.; Baddela, V.S.; Chand, S.; Baithalu, R.K.; Singh, D.; Onteru, S.K. Salivary miR-16, miR-191 and miR-223: Intuitive indicators of dominant ovarian follicles in buffaloes. Mol. Genet. Genom. 2017, 292, 935–953. [Google Scholar] [CrossRef]
  94. Kasimanickam, V.; Kastelic, J. Circulating cell-free mature microRNAs and their target gene prediction in bovine metritis. Sci. Rep. 2016, 6, 29509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Peñagaricano, F.; Souza, A.H.; Carvalho, P.D.; Driver, A.M.; Gambra, R.; Kropp, J.; Hackbart, K.S.; Luchini, D.; Shaver, R.D.; Wiltbank, M.C.; et al. Effect of Maternal Methionine Supplementation on the Transcriptome of Bovine Preimplantation Embryos. PLoS ONE 2013, 8, e72302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Ly, L.; Chan, D.; Trasler, J.M. Developmental windows of susceptibility for epigenetic inheritance through the male germline. Semin. Cell Dev. Biol. 2015, 43, 96–105. [Google Scholar] [CrossRef] [PubMed]
  97. Terashima, M.; Barbour, S.; Ren, J.; Yu, W.; Han, Y.; Muegge, K. Effect of high fat diet on paternal sperm histone distribution and male offspring liver gene expression. Epigenetics 2015, 10, 861–871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Stewart, K.R.; Veselovska, L.; Kelsey, G. Establishment and functions of DNA methylation in the germline. Epigenomics 2016, 8, 1399–1413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Butler, W.R.; Smith, R.D. Relationships between energy balance and postpartum reproductive function in dairy cattle. J. Dairy Sci. 1989, 72, 767–783. [Google Scholar] [CrossRef]
  100. Butler, W.R. Review: Effect of Protein Nutrition on Ovarian and Uterine Physiology in Dairy Cattle. J. Dairy Sci. 1998, 81, 2533–2539. [Google Scholar] [CrossRef]
  101. Gath, V.P.; Crowe, M.A.; O’Callaghan, D.; Boland, M.P.; Duffy, P.; Lonergan, P.; Mulligan, F.J. Effects of diet type on establishment of pregnancy and embryo development in beef heifers. Anim. Reprod. Sci. 2012, 133, 139–145. [Google Scholar] [CrossRef]
  102. Byrne, C.J.; Fair, S.; English, A.M.; Cirot, M.; Staub, C.; Lonergan, P.; Kenny, D.A. Plane of nutrition before and after 6 months of age in Holstein-Friesian bulls: I. Effects on performance, body composition, age at puberty, and postpubertal semen production. J. Dairy Sci. 2018, 101, 3447–3459. [Google Scholar] [CrossRef] [Green Version]
  103. Byrne, C.J.; Fair, S.; English, A.-M.; Urh, C.; Sauerwein, H.; Crowe, M.A.; Lonergan, P.; Kenny, D.A. Plane of nutrition before and after 6 months of age in Holstein-Friesian bulls: II. Effects on metabolic and reproductive endocrinology and identification of physiological markers of puberty and sexual maturation. J. Dairy Sci. 2018, 101, 3460–3475. [Google Scholar] [CrossRef]
  104. Arangasamy, A.; Sharma, R.B.; Hemalatha, K.; Krishnaiah, M.V.; Selvaraju, S.; Rani, G.P.; Binsila, B.K.; Soren, N.M.; Reddy, I.J.; Ravindra, J.P. Relationship of organic mineral supplementation and spermatozoa/white blood cells mRNA in goats. Anim. Reprod. Sci. 2018, 197, 296–304. [Google Scholar] [CrossRef] [PubMed]
  105. Zheng, W.; Feng, N.; Wang, Y.; Noll, L.; Xu, S.; Liu, X.; Lu, N.; Zou, H.; Gu, J.; Yuan, Y.; et al. Effects of zearalenone and its derivatives on the synthesis and secretion of mammalian sex steroid hormones: A review. Food Chem. Toxicol. 2019, 126, 262–276. [Google Scholar] [CrossRef] [PubMed]
  106. Santos, J.E.P.; Bilby, T.R.; Thatcher, W.W.; Staples, C.R.; Silvestre, F.T. Long Chain Fatty Acids of Diet as Factors Influencing Reproduction in Cattle. Reprod. Domest. Anim. 2008, 43, 23–30. [Google Scholar] [CrossRef] [PubMed]
  107. Schoenberg, K.M.; Overton, T.R. Effects of plane of nutrition and 2,4-thiazolidinedione on insulin responses and adipose tissue gene expression in dairy cattle during late gestation. J. Dairy Sci. 2011, 94, 6021–6035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Schoenberg, K.M.; Perfield, K.L.; Farney, J.K.; Bradford, B.J.; Boisclair, Y.R.; Overton, T.R. Effects of prepartum 2,4-thiazolidinedione on insulin sensitivity, plasma concentrations of tumor necrosis factor-α and leptin, and adipose tissue gene expression. J. Dairy Sci. 2011, 94, 5523–5532. [Google Scholar] [CrossRef] [PubMed]
  109. Gulliver, C.E.; Friend, M.A.; King, B.J.; Clayton, E.H. The role of omega-3 polyunsaturated fatty acids in reproduction of sheep and cattle. Anim. Reprod. Sci. 2012, 131, 9–22. [Google Scholar] [CrossRef]
  110. Leroy, J.; Sturmey, R.G.; Van Hoeck, V.; De Bie, J.; McKeegan, P.J.; Bols, P.E.J. Dietary Fat Supplementation and the Consequences for Oocyte and Embryo Quality: Hype or Significant Benefit for Dairy Cow Reproduction? Reprod. Domest. Anim. 2014, 49, 353–361. [Google Scholar] [CrossRef]
  111. Hosseini, A.; Tariq, M.R.; Trindade da Rosa, F.; Kesser, J.; Iqbal, Z.; Mora, O.; Sauerwein, H.; Drackley, J.K.; Trevisi, E.; Loor, J.J. Insulin Sensitivity in Adipose and Skeletal Muscle Tissue of Dairy Cows in Response to Dietary Energy Level and 2,4-Thiazolidinedione (TZD). PLoS ONE 2015, 10, e0142633. [Google Scholar] [CrossRef]
  112. Fan, Y.X.; Wang, Z.; Ren, C.F.; Ma, T.W.; Deng, K.P.; Feng, X.; Li, F.Z.; Wang, F.; Zhang, Y.L. Effect of dietary energy restriction and subsequent compensatory feeding on testicular transcriptome in developing rams. Theriogenology 2018, 119, 198–207. [Google Scholar] [CrossRef]
  113. Qu, Y.-H.; Jian, L.-Y.; Ce, L.; Ma, Y.; Xu, C.-C.; Gao, Y.-F.; Machaty, Z.; Luo, H.-L. Identification of candidate genes in regulation of spermatogenesis in sheep testis following dietary vitamin E supplementation. Anim. Reprod. Sci. 2019, 205, 52–61. [Google Scholar] [CrossRef]
  114. Yadav, D.; Singh, A.K.; Kumar, B.; Mahla, A.S.; Singh, S.K.; Patra, M.K.; Kumar, H.; Kumar, S.; Tyagi, B.; Verma, M.R.; et al. Effect of n-3 PUFA-rich fish oil supplementation during late gestation on kidding, uterine involution and resumption of follicular activity in goat. Reprod. Domest. Anim. 2019, 54, 1651–1659. [Google Scholar] [CrossRef] [PubMed]
  115. Ho, E.; Dashwood, R.H. Dietary Manipulation of Histone Structure and Function. World Rev. Nutr. Diet. 2010, 101, 95–102. [Google Scholar] [PubMed]
  116. Begum, G.; Stevens, A.; Smith, E.B.; Connor, K.; Challis, J.R.G.; Bloomfield, F.; White, A. Epigenetic changes in fetal hypothalamic energy regulating pathways are associated with maternal undernutrition and twinning. FASEB J. 2012, 26, 1694–1703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Sinclair, K.D.; Allegrucci, C.; Singh, R.; Gardner, D.S.; Sebastian, S.; Bispham, J.; Thurston, A.; Huntley, J.F.; Rees, W.D.; Maloney, C.A.; et al. {DNA} methylation, insulin resistance, and blood pressure in offspring determined by maternal periconceptional B vitamin and methionine status. Proc. Natl. Acad. Sci. USA 2007, 104, 19351–19356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Osorio, J.S.; Jacometo, C.B.; Zhou, Z.; Luchini, D.; Cardoso, F.C.; Loor, J.J. Hepatic global DNA and peroxisome proliferator-activated receptor alpha promoter methylation are altered in peripartal dairy cows fed rumen-protected methionine. J. Dairy Sci. 2016, 99, 234–244. [Google Scholar] [CrossRef] [Green Version]
  119. Chadio, S.; Kotsampasi, B.; Taka, S.; Liandris, E.; Papadopoulos, N.; Plakokefalos, E. Epigenetic changes of hepatic glucocorticoid receptor in sheep male offspring undernourished in utero. Reprod. Fertil. Dev. 2017, 29, 1995. [Google Scholar] [CrossRef]
  120. Zhang, S.; Rattanatray, L.; McMillen, I.C.; Suter, C.M.; Morrison, J.L. Periconceptional nutrition and the early programming of a life of obesity or adversity. Prog. Biophys. Mol. Biol. 2011, 106, 307–314. [Google Scholar] [CrossRef]
  121. Batistel, F.; Alharthi, A.S.; Yambao, R.R.C.; Elolimy, A.A.; Pan, Y.-X.; Parys, C.; Loor, J.J. Methionine supply during late-gestation triggers offspring sex-specific divergent changes in metabolic and epigenetic signatures in bovine placenta. J. Nutr. 2019, 149, 6–17. [Google Scholar] [CrossRef]
  122. Yan, X.; Zhu, M.-J.; Dodson, M.V.; Du, M. Developmental Programming of Fetal Skeletal Muscle and Adipose Tissue Development. J. Genom. 2013, 1, 29–38. [Google Scholar] [CrossRef] [Green Version]
  123. Du, M.; Tong, J.; Zhao, J.; Underwood, K.R.; Zhu, M.; Ford, S.P.; Nathanielsz, P.W. Fetal programming of skeletal muscle development in ruminant animals. J. Anim. Sci. 2010, 88, E51–E60. [Google Scholar] [CrossRef] [Green Version]
  124. Jacometo, C.B.; Zhou, Z.; Luchini, D.; Corrêa, M.N.; Loor, J.J. Maternal supplementation with rumen-protected methionine increases prepartal plasma methionine concentration and alters hepatic mRNA abundance of 1-carbon, methionine, and transsulfuration pathways in neonatal Holstein calves. J. Dairy Sci. 2017, 100, 3209–3219. [Google Scholar] [CrossRef] [PubMed]
  125. Awadeh, F.T.; Kincaid, R.L.; Johnson, K.A. Effect of Level and Source of Dietary Selenium on Concentrations of Thyroid Hormones and Immunoglobulins in Beef Cows and Calves. J. Anim. Sci. 1998, 76, 1204–1215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Wang, X.; Lan, X.; Radunz, A.E.; Khatib, H. Maternal nutrition during pregnancy is associated with differential expression of imprinted genes and DNA methyltranfereases in muscle of beef cattle offspring. J. Anim. Sci. 2015, 93, 35–40. [Google Scholar] [CrossRef] [Green Version]
  127. Liu, S.; Lei, J.; Hancock, S.; Scanlan, V.; Broomfield, S.; Currie, A.; Thompson, A. Lamb survival, glutathione redox state and immune function of neonates and lambs from periparturient Merino ewes supplemented with rumen-protected methionine. Arch. Anim. Nutr. 2016, 70, 389–401. [Google Scholar] [CrossRef]
  128. Batistel, F.; Alharthi, A.S.M.; Wang, L.; Parys, C.; Pan, Y.-X.; Cardoso, F.C.; Loor, J.J. Placentome Nutrient Transporters and Mammalian Target of Rapamycin Signaling Proteins Are Altered by the Methionine Supply during Late Gestation in Dairy Cows and Are Associated with Newborn Birth Weight. J. Nutr. 2017, 147, 1640–1647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Moisá, S.J.; Shike, D.W.; Shoup, L.; Rodriguez-Zas, S.L.; Loor, J.J. Maternal Plane of Nutrition during Late Gestation and Weaning Age Alter Angus × Simmental Offspring Longissimus Muscle Transcriptome and Intramuscular Fat. PLoS ONE 2015, 10, e0131478. [Google Scholar] [CrossRef] [PubMed]
  130. Moisá, S.J.; Shike, D.W.; Shoup, L.; Loor, J.J. Maternal plane of nutrition during late-gestation and weaning age alter steer calf longissimus muscle adipogenic microRNA and target gene expression. Lipids 2016, 51, 123–138. [Google Scholar] [CrossRef]
  131. Chavatte-Palmer, P.; Velazquez, M.A.; Jammes, H.; Duranthon, V. Review: Epigenetics, developmental programming and nutrition in herbivores. Animal 2018, 12, S363–S371. [Google Scholar] [CrossRef] [Green Version]
  132. Godden, S. Colostrum Management for Dairy Calves. Vet. Clin. North Am. Food Anim. Pract. 2008, 24, 19–39. [Google Scholar] [CrossRef]
  133. Baumrucker, C.; Hadsell, D.L.; Blum, J.W. Effects of dietary insulin-like growth factor I on growth and insulin-like growth factor receptors in neonatal calf intestine. J. Anim. Sci. 1994, 72, 428–433. [Google Scholar] [CrossRef]
  134. Blättler, U.; Hammon, H.M.; Morel, C.; Philipona, C.; Rauprich, A.; Romé, V.; Le Huërou-Luron, I.; Guilloteau, P.; Blum, J.W. Feeding Colostrum, Its Composition and Feeding Duration Variably Modify Proliferation and Morphology of the Intestine and Digestive Enzyme Activities of Neonatal Calves. J. Nutr. 2001, 131, 1256–1263. [Google Scholar] [CrossRef] [PubMed]
  135. Liang, G.; Malmuthuge, N.; Bao, H.; Stothard, P.; Griebel, P.J. Transcriptome analysis reveals regional and temporal differences in mucosal immune system development in the small intestine of neonatal calves. BMC Genom. 2016, 17, 602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Fischer, A.J.; Song, Y.; He, Z.; Haines, D.M.; Guan, L.L.; Steele, M.A. Effect of delaying colostrum feeding on passive transfer and intestinal bacterial colonization in neonatal male Holstein calves. J. Dairy Sci. 2018, 101, 3099–3109. [Google Scholar] [CrossRef] [Green Version]
  137. He, Z.; Fischer, A.; Song, Y.; Steele, M. Genome wide transcriptome analysis provides bases on colonic mucosal immune system development affected by colostrum feeding strategies in neonatal calves. BMC Genom. 2018, 19, 635. [Google Scholar] [CrossRef] [Green Version]
  138. Zhao, X.W.; Qi, Y.X.; Huang, D.W.; Pan, X.C.; Cheng, G.L.; Zhao, H.L.; Yang, Y.X. Changes in serum metabolites in response to ingested colostrum and milk in neonatal calves, measured by nuclear magnetic resonance-based metabolomics analysis. J. Dairy Sci. 2018, 101, 7168–7181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Li, Z.; Liu, H.; Jin, X.; Lo, L.; Liu, J. Expression profiles of microRNAs from lactating and non-lactating bovine mammary glands and identification of miRNA related to lactation. BMC Genom. 2012, 13, 731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Kosaka, N.; Izumi, H.; Sekine, K.; Ochiya, T. microRNA as a new immune-regulatory agent in breast milk. Silence 2010, 1, 7. [Google Scholar] [CrossRef] [Green Version]
  141. Van Hese, I.; Goossens, K.; Vandaele, L.; Opsomer, G. Invited review: MicroRNAs in bovine colostrum—Focus on their origin and potential health benefits for the calf. J. Dairy Sci. 2020, 103, 1–15. [Google Scholar] [CrossRef] [Green Version]
  142. Li, R.; Beaudoin, F.; Ammah, A.A.; Bissonnette, N.; Benchaar, C.; Zhao, X.; Lei, C.; Ibeagha-Awemu, E.M. Deep sequencing shows microRNA involvement in bovine mammary gland adaptation to diets supplemented with linseed oil or safflower oil. BMC Genom. 2015, 16, 884. [Google Scholar] [CrossRef] [Green Version]
  143. Malmuthuge, N.; Li, M.; Fries, P.; Griebel, P.J. Regional and age dependent changes in gene expression of Toll-like receptors and key antimicrobial defence molecules throughout the gastrointestinal tract of dairy calves. Vet. Immunol. Immunopathol. 2012, 146, 18–26. [Google Scholar] [CrossRef]
  144. Malmuthuge, N.; Liang, G. Regulation of rumen development in neonatal ruminants through microbial metagenomes and host transcriptomes. Genome Biol. 2019, 20, 172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Malmuthuge, N.; Chen, Y.; Liang, G.; Goonewardene, L.A.; Guan, L.L. Heat-treated colostrum feeding promotes beneficial bacteria colonization in the small intestine of neonatal calves. J. Dairy Sci. 2015, 98, 8044–8053. [Google Scholar] [CrossRef] [PubMed]
  146. Malmuthuge, N.; Griebel, P.J.; Guan, L.L. The Gut Microbiome and Its Potential Role in the Development and Function of Newborn Calf Gastrointestinal Tract. Front. Vet. Sci. 2015, 2, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Oikonomou, G.; Teixeira, A.G.V.; Foditsch, C.; Bicalho, M.L.; Machado, V.S.; Bicalho, R.C. Fecal Microbial Diversity in Pre-Weaned Dairy Calves as Described by Pyrosequencing of Metagenomic 16S rDNA. Associations of Faecalibacterium Species with Health and Growth. PLoS ONE 2013, 8, e63157. [Google Scholar] [CrossRef]
  148. Jing, X.; Wang, W.; Degen, A.; Guo, Y.; Kang, J.; Liu, P.; Ding, L.; Shang, Z.; Fievez, V.; Zhou, J. Tibetan sheep have a high capacity to absorb and to regulate metabolism of SCFA in the rumen epithelium to adapt to low energy intake. Br. J. Nutr. 2020, 123, 721–736. [Google Scholar] [CrossRef]
  149. Flaga, J.; Górka, P.; Zabielski, R.; Kowalski, Z.M. Differences in monocarboxylic acid transporter type 1 expression in rumen epithelium of newborn calves due to age and milk or milk replacer feeding. J. Anim. Physiol. Anim. Nutr. 2015, 99, 521–530. [Google Scholar] [CrossRef]
  150. Kuzinski, J.; Röntgen, M. The mRNA and protein expression of ruminal MCT1 is increased by feeding a mixed hay/concentrate diet compared with hay ad libitum diet. Arch. Anim. Breed. 2011, 54, 280–286. [Google Scholar] [CrossRef]
  151. Moisá, S.J.; Shike, D.W.; Faulkner, D.B.; Meteer, W.T.; Keisler, D.; Loor, J.J. Central role of the PPAR$γ$ gene network in coordinating beef cattle intramuscular adipogenesis in response to weaning age and nutrition. Gene Regul. Syst. Bio. 2014, 8, 17–32. [Google Scholar]
  152. Naeem, A.; Drackley, J.K.; Stamey, J.; Loor, J.J. Role of metabolic and cellular proliferation genes in ruminal development in response to enhanced plane of nutrition in neonatal Holstein calves. J. Dairy Sci. 2012, 95, 1807–1820. [Google Scholar] [CrossRef] [Green Version]
  153. Soberon, F.; Raffrenato, E.; Everett, R.W.; Van Amburgh, M.E. Preweaning milk replacer intake and effects on long-term productivity of dairy calves. J. Dairy Sci. 2012, 95, 783–793. [Google Scholar] [CrossRef] [Green Version]
  154. Soberon, F.; Van Amburgh, M.E. The effect of nutrient intake from milk or milk replacer of preweaned dairy calves on lactation milk yield as adults: A meta-analysis of current data. J. Anim. Sci. 2013, 91, 706–712. [Google Scholar] [CrossRef] [PubMed]
  155. Fischer, A.J.; Villot, C.; van Niekerk, J.K.; Yohe, T.T.; Renaud, D.L.; Steele, M.A. Invited Review: Nutritional regulation of gut function in dairy calves: From colostrum to weaning. Am. Regist. Prof. Anim. Sci. 2019, 35, 498–510. [Google Scholar] [CrossRef]
  156. Diao, Q.; Zhang, R.; Fu, T. Review of strategies to promote rumen development in calves. Animals 2019, 9, 490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Adhikari, B.; Khanal, P.; Nielsen, M.O. Impacts of pre- and postnatal nutrition on glucagon regulation and hepatic signalling in sheep. J. Endocrinol. 2018, 238, 1–12. [Google Scholar] [CrossRef] [Green Version]
  158. Alexandre, P.A.; Kogelman, L.J.A.; Santana, M.H.A.; Passarelli, D.; Pulz, L.H.; Fantinato-Neto, P.; Silva, P.L.; Leme, P.R.; Strefezzi, R.F.; Coutinho, L.L. Liver transcriptomic networks reveal main biological processes associated with feed efficiency in beef cattle. BMC Genom. 2015, 16, 1073. [Google Scholar] [CrossRef] [Green Version]
  159. Johnson, J.R.; Carstens, G.E.; Krueger, W.K.; Lancaster, P.A.; Brown, E.G.; Tedeschi, L.O.; Anderson, R.C.; Johnson, K.A.; Brosh, A. Associations between residual feed intake and apparent nutrient digestibility, in vitro methane-producing activity, and volatile fatty acid concentrations in growing beef cattle. J. Anim. Sci. 2019, 97, 3550–3561. [Google Scholar] [CrossRef]
  160. Nkrumah, J.D.; Okine, E.K.; Mathison, G.W.; Schmid, K.; Li, C.; Basarab, J.A.; Price, M.A.; Wang, Z.; Moore, S.S. Relationships of feedlot feed efficiency, performance, and feeding behavior with metabolic rate, methane production, and energy partitioning in beef cattle. J. Anim. Sci. 2006, 84, 145–153. [Google Scholar] [CrossRef]
  161. Okine, E.K.; Basarab, J.A.; Baron, V.; Price, M.A. Net feed efficiency in young growing cattle: III. Relationship to methane and manure production. In Proceedings of the Presentations and Posters, Agricultural Institute of Canada, 2001 Conference’, University of Guelph, Guelph, ON, Canada, 29 July–2 August 2001. [Google Scholar]
  162. Schurmann, B.L.; Walpole, M.E.; Górka, P.; Ching, J.C.H.; Loewen, M.E.; Penner, G.B. Integrative and Translational Physiology: Inflammation, Immunity, and Organ System Physiology: Short-term adaptation of the ruminal epithelium involves abrupt changes in sodium and short-chain fatty acid transport. Am. J. Physiol. Integr. Comp. Physiol. 2014, 307, R802. [Google Scholar] [CrossRef] [Green Version]
  163. Gholizadeh, M.; Fayazi, J.; Asgari, Y.; Zali, H.; Kaderali, L. Reconstruction and Analysis of Cattle Metabolic Networks in Normal and Acidosis Rumen Tissue. Animals 2020, 10, 469. [Google Scholar] [CrossRef] [Green Version]
  164. Lu, Z.; Yao, L.; Jiang, Z.; Aschenbach, J.R.; Martens, H.; Shen, Z. Acidic pH and short-chain fatty acids activate Na+ transport but differentially modulate expression of Na+/H+ exchanger isoforms 1, 2, and 3 in omasal epithelium. J. Dairy Sci. 2016, 99, 733–745. [Google Scholar] [CrossRef] [Green Version]
  165. Yang, W.; Shen, Z.; Martens, H. An energy-rich diet enhances expression of Na+/H+ exchanger isoform 1 and 3 messenger RNA in rumen epithelium of goat. J. Anim. Sci. 2012, 90, 307–317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Yan, L.; Shen, Z.; Lu, Z. Increases in the expression of Na+ /H+ exchanger 1 and 3 are associated with insulin signalling in the ruminal epithelium. J. Anim. Physiol. Anim. Nutr. 2017, 102, e569–e577. [Google Scholar] [CrossRef] [PubMed]
  167. Penner, G.B.; Steele, M.A.; Aschenbach, J.R.; McBride, B.W. Ruminant Nutrition Symposium: Molecular adaptation of ruminal epithelia to highly fermentable diets. J. Anim. Sci. 2011, 89, 1108–1119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Stumpff, F. A look at the smelly side of physiology: Transport of short chain fatty acids. Pflug. Arch. 2018, 470, 571–598. [Google Scholar] [CrossRef] [PubMed]
  169. Arroyo, J.M.; Hosseini, A.; Zhou, Z.; Alharthi, A.; Trevisi, E.; Osorio, J.S.; Loor, J.J. Reticulo-rumen mass, epithelium gene expression, and systemic biomarkers of metabolism and inflammation in Holstein dairy cows fed a high-energy diet. J. Dairy Sci. 2017, 100, 9352–9360. [Google Scholar] [CrossRef]
  170. Baaske, L.; Gäbel, G.; Dengler, F. Ruminal epithelium: A checkpoint for cattle health. J. Dairy Res. 2020, 87, 322–329. [Google Scholar] [CrossRef]
  171. Baaske, L.; Masur, F.; Dengler, F.; Rackwitz, R.; Kaiser, B.; Pfannkuche, H.; Gäbel, G. Possible influence of free fatty acid receptors on pH regulation in the ruminal epithelium of sheep. J. Anim. Physiol. Anim. Nutr. 2020, 104, 776–789. [Google Scholar] [CrossRef] [Green Version]
  172. Herve, L.; Quesnel, H.; Veron, M.; Portanguen, J.; Gross, J.J.; Bruckmaier, R.M.; Boutinaud, M. Milk yield loss in response to feed restriction is associated with mammary epithelial cell exfoliation in dairy cows. J. Dairy Sci. 2019, 102, 2670–2685. [Google Scholar] [CrossRef] [Green Version]
  173. Meale, S.J.; Romao, J.M.; He, M.L.; Chaves, A.V.; McAllister, T.A.; Guan, L.L. Effect of diet on microRNA expression in ovine subcutaneous and visceral adipose tissues. J. Anim. Sci. 2014, 92, 3328–3337. [Google Scholar] [CrossRef] [Green Version]
  174. Tsai, C.Y.; Rezamand, P.; Loucks, W.I.; Scholte, C.M.; Doumit, M.E. The effect of dietary fat on fatty acid composition, gene expression and vitamin status in pre-ruminant calves. Anim. Feed Sci. Technol. 2017, 229, 32–42. [Google Scholar] [CrossRef]
  175. Loor, J.J.; Dann, H.M.; Janovick Guretzky, N.A.; Everts, R.E.; Oliveira, R.; Green, C.A.; Litherland, N.B.; Rodriguez-Zas, S.L.; Lewin, H.A.; Drackley, J.K. Plane of nutrition prepartum alters hepatic gene expression and function in dairy cows as assessed by longitudinal transcript and metabolic profiling. Physiol. Genom. 2006, 27, 29–41. [Google Scholar] [CrossRef] [PubMed]
  176. Palmquist, D.L.; Jenkins, T.C. Fat in lactation rations. J. Dairy Sci. 1980, 63, 1–14. [Google Scholar] [CrossRef]
  177. Diskin, M.G.; Morris, D.G. Embryonic and Early Foetal Losses in Cattle and Other Ruminants. Reprod. Domest. Anim. 2008, 43, 260–267. [Google Scholar] [CrossRef] [PubMed]
  178. Santos, A.; Giráldez, F.J.; Valdés, C.; Trevisi, E.; Lucini, L.; Frutos, J.; Andrés, S. Milk replacer restriction during early life impairs the live body weight and progesterone patterns of ewe lambs during the replacement period. J. Dairy Sci. 2018, 101, 8021–8031. [Google Scholar] [CrossRef]
  179. Elolimy, A.A.; Abdelmegeid, M.K.; McCann, J.C.; Shike, D.W.; Loor, J.J. Residual feed intake in beef cattle and its association with carcass traits, ruminal solid-fraction bacteria, and epithelium gene expression. J. Anim. Sci. Biotechnol. 2018, 9, 67. [Google Scholar] [CrossRef]
  180. De Oliveira, P.S.N.; Coutinho, L.L.; Tizioto, P.C.; Cesar, A.S.M.; de Oliveira, G.B.; Diniz, W.J.d.S.; De Lima, A.O.; Reecy, J.M.; Mourão, G.B.; Zerlotini, A.; et al. An integrative transcriptome analysis indicates regulatory mRNA-miRNA networks for residual feed intake in Nelore cattle. Sci. Rep. 2018, 8, 17072. [Google Scholar] [CrossRef]
  181. Liu, L.; Sun, D.; Mao, S.; Zhu, W.; Liu, J. Infusion of sodium butyrate promotes rumen papillae growth and enhances expression of genes related to rumen epithelial VFA uptake and metabolism in neonatal twin lambs. J. Anim. Sci. 2019, 97, 909–921. [Google Scholar] [CrossRef]
  182. Zhan, K.; Yang, T.Y.; Chen, Y.; Jiang, M.C.; Zhao, G.Q. Propionate enhances the expression of key genes involved in the gluconeogenic pathway in bovine intestinal epithelial cells. J. Dairy Sci. 2020, 103, 5514–5524. [Google Scholar] [CrossRef]
  183. Aschenbach, J.R.; Zebeli, Q.; Patra, A.K.; Greco, G.; Amasheh, S.; Penner, G.B. Symposium review: The importance of the ruminal epithelial barrier for a healthy and productive cow. J. Dairy Sci. 2019, 102, 1866–1882. [Google Scholar] [CrossRef] [Green Version]
  184. Dengler, F.; Rackwitz, R.; Benesch, F.; Pfannkuche, H.; Gäbel, G. Both butyrate incubation and hypoxia upregulate genes involved in the ruminal transport of SCFA and their metabolites. J. Anim. Physiol. Anim. Nutr. 2015, 99, 379–390. [Google Scholar] [CrossRef]
  185. Yan, L.; Zhang, B.; Shen, Z. Dietary modulation of the expression of genes involved in short-chain fatty acid absorption in the rumen epithelium is related to short-chain fatty acid concentration and pH in the rumen of goats. J. Dairy Sci. 2014, 97, 5668–5675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Wang, J.-F.; Fu, S.-P.; Li, S.-N.; Hu, Z.-M.; Xue, W.-J.; Li, Z.-Q.; Huang, B.-X.; Lv, Q.-K.; Liu, J.-X.; Wang, W. Short-Chain Fatty Acids Inhibit Growth Hormone and Prolactin Gene Transcription via cAMP/PKA/CREB Signaling Pathway in Dairy Cow Anterior Pituitary Cells. Int. J. Mol. Sci. 2013, 14, 21474–21488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Kasper, C.; Ribeiro, D.; de Almeida, A.M.; Larzul, C.; Liaubet, L.; Murani, E. Omics application in animal science—a special emphasis on stress response and damaging behaviour in pigs. Genes 2020, 11, 920. [Google Scholar] [CrossRef] [PubMed]
  188. Zhu, H.L.; Zhao, X.W.; Han, R.W.; DU, Q.J.; Qi, Y.X.; Jiang, H.N.; Huang, D.W.; Yang, Y.X. Changes in bacterial community and expression of genes involved in intestinal innate immunity in the jejunum of newborn lambs during the first 24 hours of life. J. Dairy Sci. 2021, 104, 9263–9275. [Google Scholar] [CrossRef] [PubMed]
  189. Osorio, J. Gut health, stress, and immunity in neonatal dairy calves: The host side of host-pathogen interactions. J. Anim. Sci. Biotechnol. 2020, 11, 105. [Google Scholar] [CrossRef]
Figure 1. Potential of different dietary components and characteristics of molecular changes in a ruminant model.
Figure 1. Potential of different dietary components and characteristics of molecular changes in a ruminant model.
Animals 12 00997 g001
Figure 2. Nutrition–gene interaction pathway in ruminants.
Figure 2. Nutrition–gene interaction pathway in ruminants.
Animals 12 00997 g002
Figure 3. Animal factors that can change the feedomics imprint.
Figure 3. Animal factors that can change the feedomics imprint.
Animals 12 00997 g003
Table 1. Dietary characteristics and components that have an epigenetic effect.
Table 1. Dietary characteristics and components that have an epigenetic effect.
FactorActionAnimal TypeReference
Maternal protein insufficiencyDNA methylationSheep[116]
Vitamin b12, folate, and methionine deficiencyDNA methylationSheep[117]
Rumen-protectedDNA methylationCows[118]
methionine
Maternal undernutritionDNA methylationSheep[119]
Maternal overnutritionSex-specific DNA methylationSheep[120]
Methionine supplySex-specific DNA methylationCows[121]
UndernutritionMicroRNAsCows[122,123]
Histone modifications
DNA methylation
Rumen-protected methionineDNA methylationCows[124]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abdelrahman, M.; Wang, W.; Shaukat, A.; Kulyar, M.F.-e.-A.; Lv, H.; Abulaiti, A.; Yao, Z.; Ahmad, M.J.; Liang, A.; Yang, L. Nutritional Modulation, Gut, and Omics Crosstalk in Ruminants. Animals 2022, 12, 997. https://doi.org/10.3390/ani12080997

AMA Style

Abdelrahman M, Wang W, Shaukat A, Kulyar MF-e-A, Lv H, Abulaiti A, Yao Z, Ahmad MJ, Liang A, Yang L. Nutritional Modulation, Gut, and Omics Crosstalk in Ruminants. Animals. 2022; 12(8):997. https://doi.org/10.3390/ani12080997

Chicago/Turabian Style

Abdelrahman, Mohamed, Wei Wang, Aftab Shaukat, Muhammad Fakhar-e-Alam Kulyar, Haimiao Lv, Adili Abulaiti, Zhiqiu Yao, Muhammad Jamil Ahmad, Aixin Liang, and Liguo Yang. 2022. "Nutritional Modulation, Gut, and Omics Crosstalk in Ruminants" Animals 12, no. 8: 997. https://doi.org/10.3390/ani12080997

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