5,283
Views
116
CrossRef citations to date
0
Altmetric
Reviews

Intermuscular and intramuscular adipose tissues: Bad vs. good adipose tissues

, , , &
Pages 242-255 | Received 06 Dec 2013, Accepted 14 Mar 2014, Published online: 30 Oct 2014

Abstract

Human studies of the influence of aging and other factors on intermuscular fat (INTMF) were reviewed. Intermuscular fat increased with weight loss, weight gain, or with no weight change with age in humans. An increase in INTMF represents a similar threat to type 2 diabetes and insulin resistance as does visceral adipose tissue (VAT). Studies of INTMF in animals covered topics such as quantitative deposition and genetic relationships with other fat depots. The relationship between leanness and higher proportions of INTMF fat in pigs was not observed in human studies and was not corroborated by other pig studies. In humans, changes in muscle mass, strength and quality are associated with INTMF accretion with aging. Gene expression profiling and intrinsic methylation differences in pigs demonstrated that INTMF and VAT are primarily associated with inflammatory and immune processes. It seems that in the pig and humans, INTMF and VAT share a similar pattern of distribution and a similar association of components dictating insulin sensitivity. Studies on intramuscular (IM) adipocyte development in meat animals were reviewed. Gene expression analysis and genetic analysis have identified candidate genes involved in IM adipocyte development. Intramuscular (IM) adipocyte development in human muscle is only seen during aging and some pathological circumstance. Several genetic links between human and meat animal adipogenesis have been identified. In pigs, the Lipin1 and Lipin 2 gene have strong genetic effects on IM accumulation. Lipin1 deficiency results in immature adipocyte development in human lipodystrophy. In humans, overexpression of Perilipin 2 (PLIN2) facilitates intramyocellular lipid accretion whereas in pigs PLIN2 gene expression is associated with IM deposition. Lipins and perilipins may influence intramuscular lipid regardless of species.

Introduction

By their very nature, problems associated with obesity, diabetes, and metabolic syndrome are being itemized, categorized, and legitimized by the amount of research efforts directed in elucidating resolutions to adverse symptoms associated with these and other metabolic dysfunctions.Citation1-5 Type I diabetes is known, in part, to be caused by a lack of insulin due to faulty production by β cells of the pancreas, autoantibody removal of insulin in circulation causing sequestration/inactivation, and inhibition or failure of insulin to bind to somatic cells such as skeletal muscle due to a decrease in insulin receptor number or an increase in anti-receptor antibodies binding to the insulin receptor.Citation6,7 Type II diabetes, metabolic syndrome, and obesity, however, are highly researched but not thoroughly resolved, and result in health issues that are becoming elevated to world-wide crises.Citation1-5 Resident in these issues are research efforts to determine cellular and molecular control of depot-specific adipocytes, interaction, and association of adipocytes to other somatic cells, disproportionate propensities of depot-specific adipocytes to produce adipokines that function to control numerous types of body cells, and adipocyte propensity to (seemingly) disassociate themselves from normal regulatory function/control during metabolic dysfunctions such as obesity—likely as a consequence of hypoxia in the adipose depot.Citation2-4,Citation8-12

Adipocytes found in different anatomical adipose depots express divergent physiologies throughout one's lifespan.Citation9,10,13,14 The five traditional adipose depots in animals and humans are subcutaneous (SQF), visceral (VAT), intermuscular (INTMF), intramuscular (IM), and bone adipose depots. The SQF depot develops and fills with lipid first, followed by the VAT, INTMF, and IM adipose depots, and considerable research efforts are focused on depot differences in the structure, function and/or regulation of all cells contained within the depots.Citation4,9,10,14,15 In terms of anatomical location, considerable interest has been shown in the abdominal adipocytes of humans due to the possible production of adipokines that appear to function in physiologies such as control of satiety.Citation8-10 Issues associated with adipose depot include disruption of normal function of organs due to infiltration with lipid-filled adipocytes—for example, non-alcoholic fatty livers are not capable of functioning as properly as normal livers, and is likely a precursor of fatty infiltration into other organs/tissues.Citation9,16 An interest of this paper is on the specific differences in INTMF vs. IM adipose tissues.

Adipocytes associated with skeletal muscle include the INTMF adipose depot, which associates itself with available spaces in between skeletal muscles, and the IM adipose depot which includes all adipocytes interjecting themselves between and among viable skeletal muscle fibers in the skeletal muscle bed.Citation9,10,17,18 In meat animals, the IM adipose depot is termed marbling fat.Citation9,10,Citation17-19 Early development of these two adipose depots occurs during embryogenesis, becomes distinct and committed in the early part of the third trimester in utero, becomes evident during rapid periods of postnatal growth, and slows in the INTMF adipose depot while increasing in the IM adipose depot during late adolescence.Citation17,Citation19-21 Considerable knowledge exists on stem cell determination and commitment during in utero development, whereas relatively less is known about cell regulation in these two adipose depots during postnatal in vivo growth.Citation9,10,Citation18-21

A considerable number of reviews are available directing attention to preadipocyte proliferation, conversion to adipocytes and adipocyte ability to conduct lipid metabolism, produce adipokines, and become refractory to systemic signals during periods of abnormal metabolism.Citation4,6,Citation8-11,Citation14,15,Citation17-21 Much of this knowledge was generated from studies utilizing cell lines, adipose tissue stromal vascular (SV) cell cultures, or adipose tissue slices–essentially piecing together purported mechanisms of regulation and/or marker expression during cellular conversion processes.Citation10,13,15,22

Knowledge pertaining to normal regulation, dysfunctional regulation, endocrine/adipokine production, and resistance of specific adipose depots to respond to metabolic signals is growing.Citation2,Citation5-9 Integration of these concept areas to provide a mechanism with which to target for clinical remediation of the adverse symptoms of adipose tissue manifestations is lacking, however, possibly due to conflicting messages about adipose origins, the obese phenotype, its causes, and repercussions.Citation10-14 Indeed, it appears that the preponderance of evidence states that SQF adiposity supplies a measure of protection against metabolic dysregulation, whereas VAT is considered a source of dysregulation (especially insulin resistance).Citation2,Citation15-23 However, this message is further muddied by new reports that demonstrate that truncal SQF adipose tissue may actually aid in the metabolic changes that occur with obesity as this depot supplies a greater percentage of adipose tissue compared with the VAT and that the protective effects of SQ fat is limited more to the gluteal femoral SQ depot (or the so-called “pear-shape”).Citation11,15,Citation24-28 The protective action of gluteal femoral SQF seems to be further limited to the femoral subcutaneous region and not necessarily with the gluteal region in black South African women further confounding the issue with ethnicity differences.Citation11,16,25,27,Citation29-32 Indeed, even gender appears to have a profound influence on the responses by different adipose depots.Citation16,Citation32-34 One mechanism that seems universal in human adipose depot regulation of adipose tissue is insulin, although, with as much variation within one species (human) one can only imagine the differences which exist between species. Knowledge of insulin receptivity, regulation—or lack thereof—and resilience in retaining normal adipose depot physiology in the face of metabolic disturbance is vitally needed.

For hundreds of years, domestic (meat) animals have been intensely selected for a variety of carcass and quality variables. One end point of this selection pressure is IM.Citation35 Meat animals such as beef cattle are fed a grain diet in a feedlot setting prior to slaughter in order to allow the IM to form.Citation36 This feeding regimen is costly, but provides consumers a meat product with adipose cells interspersed among skeletal muscle fibers in the final meat product resulting in a meat product that is flavorful and juicy. Alternate feeding regimes have been studied with little effect on IM.Citation37-40 Such interspersed groupings of adipocytes in the musculature bed of humans are rare, and only blatantly seen during some pathological circumstance, such as lipotrophic muscular atrophy. The focus of this review is to provide detailed information about INTMF vs IM adipose tissues in humans and animals. We are particularly interested in (1) positive and negative physiological effects of these adipose depots, (2) the association of these depots with abnormal physiologies such as diabetes and metabolic syndrome, (3) gene regulation/expression and marker production by adipocytes within the two adipose depots, and (4) whether some commonalities might exist among animal and human depot-specific gene orthologs.

Molecular Regulation of Adipogenesis

Up to now, most studies on INTMF/IM adipogenesis have been conducted in mice. Due to the small size of mice, it is almost impossible to distinguish intermuscular and intramuscular fat thus all available studies of IM adipogenesis in mice refer to both INTMF and IM fat. In mice, intramuscular adipogenesis initiates later at mid to late gestation, slightly later than fetal muscle development which initiates in the mesoderm at the early fetal stage.Citation41,42 Wnt signaling plays a crucial role in activating myogenic differentiation, while inhibiting adipogenic differentiation.Citation43,44 Recently, it has been shown that brown adipocytes through involvement of PRDM16 (PR-domain-containing 16) transcription factor are derived from Myf5 positive cells which are also precursors for myogenic cells.Citation45,46 Bone morphogenetic protein 7 (BMP7) promotes brown adipogenesis through promoting PRDM16 expression in mice.Citation47 In bovine fetuses, these INTMF brown adipocytes appear to convert into white INTMF adipocytes later in life.Citation47-49 In these two studies, INTMF brown adipogenesis was not examined due to the absence of distinguishable IM fat in bovine fetal muscle. There is no evidence of INTMF/IM brown adipogenesis in pigs.

Compared with brown adipogenesis, white adipogenesis appears to have a dominant role in the formation of inter/intramuscular adipocytes.Citation48,49 Adipogenesis is briefly separated into two stages: determination and differentiation.Citation50 Recently, zinc-finger protein Zfp423 was identified as a transcriptional factor responsible for the adipogenic commitment of progenitor cells, or adipocyte determination.Citation51 The expression of Zfp423 commits progenitor cells to pre-adipocytes, which further induces peroxisome proliferator-activated receptor γ (PPARγ) expression and terminal adipogenic differentiation of cultured NIH 3T3 cells and in SQ fat.Citation51,52 It was recently observed that Xfp423 also regulates adipogenic commitment in beef stromal vascular cells.Citation53 Compared with adipogenic commitment, the differentiation of preadipocytes into mature adipocytes is much better studied. Most studies conducted in mice and 3T3-L1 cell line clearly showed that adipogenic differentiation is mediated by PPARγ and CCAAT/enhancer-binding proteins (C/EBPs).Citation54 C/EBPβ/δ, which is expressed at a very early stage of adipogenesis, triggers the expression of PPARγ, an essential and indispensable transcription factor for the late stage of adipogenesis; the co-expression of PPARγ and C/EBPα induce genes specific to adipocytes during adipogenesis.Citation55-57 As a result, cells accumulate lipid droplets and become mature adipocytes.Citation58

Animal Studies of Intermuscular Fat

In contrast to human studies, studies of INTMF fat in pigs and cattle are generally more invasive and either multiple sites or the entire INTMF fat depot is examined. Early work on the cellularity, development, and metabolism of INTMF and other adipose tissue depots in cattle, pigs, and sheep showed that INTMF cells in growing pigs, cattle, and sheep hypertrophied much less than SQF and perirenal fat cells but increased in number more than perirenal fat cells in sheep.Citation59 Furthermore, cattle and sheep INTMF fat cell cellularity were affected least by a restricted or maintenance diet than SQF and perirenal fat.Citation60,61 In a recent comparison, the porcine INTMF depot lipid content was lower and adipocytes smaller than in the SQF and perirenal depots.Citation62 Growth rates of INTMF and SQF adipose tissue are similar (relative to carcass weight) but in some locations like the belly the INTMF fat grows more rapidly.Citation62 The INTMF depot develops before 20 kg of body weight especially in lean pig breeds such as the Pietrain.Citation62 In cattle, the INTMF depot is the largest fat depot beginning at 11 mo and remains the largest until slaughter (19 mo) with the amount of SQF representing 60% of the INTMF fat mass throughout growth.Citation63 In the pig INTMF fat represents 18–13% of total fat throughout growth compared with SQF representing 78–84% of total fat.Citation62 Classical selection for increased carcass leanness, essentially based on backfat depth, appears to have been less successful for reducing INTMF accretion than for reducing subcutaneous fat in pigs, especially in lean animals. In a study of 94 pigs, seven combinations of pig genotypes and sex were serially slaughtered throughout growth between 12 and 110 kg BW and it was found that the genetically leaner animals, had a higher proportion of INTMF fat as a percent of total fat.Citation64 Among common pig breeds, the relative levels of INTMF were dependent on carcass cuts or anatomical location which were highest for the fattest pigs, i.e., Meishans and relative levels of INTMF were similar for Landrace, Large White, Pietrain, and Duroc pigs.Citation65 This discrepancy may reflect that many more pigs were used by Gispert et al.Citation65 than were examined by Kouba et al.Citation64 In a study of INTMF of four primal cuts from high and low growth rate pigs no effect of genetic selection was detected on the INTMF to SQF ratio for the primal cuts.Citation66 Feeding 5% beef tallow to pigs up to 135 kg had no influence on INTMF to SQF ratios for individual primal cuts.Citation66 The effect of a low protein diet on INTMF content and percent muscle was found to be breed specific and only significant for the Berkshire and Large White breeds.Citation67

Limited studies have examined the genetic control of INTMF fat development and composition. Genetic and phenotypic correlations between SQF and INTMF percentages were approximately 0.50 in a comparison of a lean and a fat breed of pigs.Citation62 In contrast, genetic and phenotypic correlations between INTMF cross sectional areas and IM content were smaller and ranged from 0.2 to 0.4.Citation62 Residual feed intake is strongly genetically correlated with SQF in pigs but is weakly correlated with IM.Citation68 In a later study, strong genetic correlations of residual feed intake with INTMF fat accumulations were identified at several sites in pigs indicating that reducing residual feed intake is linked to INTMF accumulations at these sites.Citation69 The genetic correlations between daily feed intake and INTMF fat accumulations at several of these sites were strong (0.60, 0.76, and 0.56 for INTMF fat at fifth to sixth thoracic vertebra, one-half body length, and last thoracic vertebra, respectively). In another study, image analysis of abdominal fat, INTMF and SQF from adipose tissue samples from Duroc pigs slaughtered at 105 kg indicated that within anatomical sites the genetic correlations between SQ and INTMF were high whereas, correlations between IM and INTMF fat were much lower.Citation70 Heritability estimates for INTMF varied between locations and were lower than SQF area heritabilities at all three anatomical sites.Citation70 In finished feed lot steers, heritability estimates for INTMF and SQF were similar (0.40, 0.42).Citation71 The absence of higher genetic correlations indicated that INTMF, SQF, and IM are regulated by different genetic means.Citation71

Molecular studies have revealed that porcine perilipin 2 (PLIN2) gene expression analyses showed a positive correlation with higher INTMF.Citation72 However, a 3′-UTR mutation PLIN2 genotype was not significantly associated with INTMF in association analysis studies.Citation72 Gene expression profiling of SQF, VAT, and INTMF adipose tissues in male and female pigs of three pig breeds with different degrees of adiposity was coupled with Gene Ontology-Biological Processes and KEGG pathway analysis.Citation73 Differentially expressed genes were identified and clustered which revealed that VAT as well as the INTMF were mainly associated with impaired inflammatory and immune response whereas SQ fat was mainly associated with metabolism modulators.Citation73 Gene modules of co-expressed genes substantiated the distinction between depots.Citation73 Furthermore, recent studies demonstrated depot dependent intrinsic DNA methylation differences in the pig and therefore demonstrated epigenetic evidence that both VAT and INTMF are associated with an impaired immune response in the pig.Citation74 In a study of miRNA transcriptomes of porcine INTMF and SQF miRNAs, INTMF transcriptomes were enriched in inflammation- and diabetes-related miRNAs compared with SQ fat transcriptomesCitation75. Moreover, functional enrichment analysis of the genes predicted to be targeted by the enriched miRNAs indicated that INTMF adipose tissue was associated with immune and inflammatory responses.Citation75 Collectively, these results indicated that SQF mainly modulates metabolism, whereas INTMF and VAT are associated with inflammatory and immune responses.Citation73-75

Human Studies of Intermuscular Fat

Recently, attention has focused on the content, localization, and composition of fat within human skeletal muscle as determinants of insulin resistance and involvement in the metabolic syndrome. Metabolic syndrome is a cluster of conditions, increased blood pressure, elevated insulin levels, excess body fat around the waist, or abnormal cholesterol levels that occur together, increasing the risk of heart disease, stroke and diabetes.Citation76 Up until now it has been thought that visceral fat or visceral obesity was responsible for the metabolic syndrome.Citation76 In humans INTMF tissue is located between muscle groups and clearly separated from SQF by a well-defined fascia. IM is located within individual muscles as observed on MRI images.Citation77 Recently, attention has focused on characteristics of fat within skeletal muscle including the intramyocellular triglycerides (IMCL), INTMF, and a small pool of adipocytes present between muscle fascicles which may be involved in insulin resistance.Citation77 Triglyceride accumulation within the muscle cell primarily accounts for IMCL.Citation78 Perivascular adipose cells and adipocytes in the intermuscular space and between muscle fascicles account for the perivascular or intermyofibrillar lipids.Citation77 With the development and adaptation of MRI and CT technology to evaluate human skeletal muscle lipid deposition many studies have examined factors that influence INTMF in humans.Citation77 With MRI or CT, INTMF is localized and studied whereas EMCL and IMCL content can be evaluated by proton magnetic resonance spectroscopy (1H MRS) or MRSI.Citation77 Additionally, IMCL can be quantified with lipid histochemistry of muscle biopsies with or without 1H MRS.Citation78

The influence of race, aging, physical activity and obesity on INTMF and muscle changes are summarized in . In a study of a cohort of a large number of older subjects, INTMF increased in those who either maintained, lost or gained weight indicating that INTMF is regulated independent of SQ fat and other fat depots ().Citation79 Another study demonstrated that physical activity can prevent losses in muscle strength and prevent increases in INTMF in the absence of changes in SQF ().Citation84 In a study of older individuals with metabolic and mobility impairments, 12 wks of resistance training decreased INTMF and increased lean tissue ().Citation81 Subjecting younger subjects to the influence of 4 wk of unilateral lower limb suspension significantly increased INTMF accretion and decreased muscle volumes with no change in SQF ().Citation80 This is a clear demonstration of how physical activity controls muscle mass and fat deposition in muscle. Healthy obese subjects (56–64 y old), diabetic subjects, and diabetic subjects with peripheral neuropathy when subjected to a 6 min walk test and a physical performance test showed muscle specific INTMF accretion ()Citation83 Evaluation of race and obesity (i.e., 1105 Caucasian and 518 Afro-Caribbean men aged 65) demonstrated that INTMF accretion was greater in African men and showed association with type 2 diabetes in both ethnic groups ().Citation85

Table 1. Influence of race, aging, reduced physical activity, and physical activity on INTMF in humans

Several of these studies indicated that changes in muscle mass, strength and quality with aging may dictate INTMF accretion in humans which was independent of total body weight or total adiposity (). In the animal studies, INTMF is regulated by genetic means distinct from the other fat depots. The relationship between leanness in pigs and higher proportions of INTMF fat was not observed in human studies and was not corroborated by several pig studies.Citation64 Gene expression profiling and intrinsic methylation differences in pigs demonstrated that INTMF and primarily VAT are primarily associated with inflammatory and immune processes and insulin resistance.Citation73,86 In humans INTMF is closely linked to VAT and is a good predictor for insulin sensitivity nearly equal to that of VAT.Citation87 Therefore, it seems that in the pig and humans, INTMF and VAT share a similar pattern in distribution and association of components dictating insulin sensitivity.

Intramuscular Adipogenesis

Overview of adipose tissue prenatal development in skeletal muscle

During the prenatal stage, skeletal muscle development mainly involves the formation of muscle fibers (i.e., myogenesis), but also the formation of IM adipocytes (adipogenesis) and fibroblasts (i.e., fibrogenesis). These cells are derived from a common pool of multipotent cells, mesenchymal progenitor cells. Based on recent discoveries, it appears that during fetal skeletal muscle development, mesenchymal multipotent cells first diverge to either myogenic or fibro/adipogenic lineages. Myogenic lineage cells further develop into muscle fibers, INTMF/IM brown adipocytes, and satellite cells, while fibro/adipogenic lineage cells develop into the stromal vascular fraction of muscle where white adipocytes, fibroblasts and resident fibro/adipogenic progenitors reside (FAPs, the counterpart of satellite cells).Citation48,53,88 These resident FAPs become largely quiescent and form the stem cell pool for later adipogenesis and fibrogenesis in mature muscle.Citation89-91 Because both myogenic and fibro/adipogenic cells are from the same pool of progenitor cells, the initial myogenic or fibro/adipogenic commitment can be considered as a competitive process, with enhancing myogenesis reducing fibro/adipogenesis, and vice versa. Indeed, both IM fat and collagen accumulation is lower while muscle mass is higher in the largest compared with the smallest littermates of pigs, indicating the shifting of myogenesis to fibro/adipogenesis in the smallest piglets.Citation92,93 Consistently, previous studies demonstrated both IM accumulation and fibrosis in fetal and offspring sheep due to maternal over-nutrition, which correlates with downregulated myogenesisCitation94-96 and suggests that IM adipogenesis and fibrogenesis are correlated due to a common cell lineage. Predictably, in genetically high marbling Wagyu cattle, both IM adipogenesis and collagen accumulation are higher than in relatively low marbling Angus cattle.Citation97

Intramuscular Adipogenesis in Livestock Species

Existing evidence points to the similarities in adipogenesis between INTMF/IM and SQF adipogenesis. The overall expression pattern of PPARγ and C/EBPα appears similar between IM and SQF SV cells, despite a temporal differences.Citation98 The expression of these markers in preadipocytes of IM is similar with those in SQF despite at  a lower level, indicating that the SV cells of IM are at a relatively earlier differentiation stage compared with that of SQF.Citation99 In addition, IM preadipocytes respond to dexamethasone treatments, similar to SQ preadipocytes but the response is greater for SQF preadipocytes in line with a recent report that the expression profiles of microRNAs between intramuscular and SQ SV cells are similar.Citation100,101 Similarly, in bovine SV cell cultures, both IM and SQF cells are responsive to dexamethasone, but the responsiveness is greater in subcutaneous cells, echoing the more advanced differentiation of SQF compared with IM SV cells.Citation102 In genetically obese pig breeds, higher expression of PPARγ in the IM preadipocytes was observed than in genetically lean breeds, indicating enhanced adipogenic differentiation.Citation86 Compared with the difference between breeds, there is more obvious difference among species. For example, the expression of both PPARγ and C/EBPα were undetectable in bovine IM preadipocytes after 4 d of adipogenic differentiation while they were expressed in swine IM SV cell cultures early on and are highly induced in mouse 3T3-L1 cell preadipocyte cultures after 4 d.Citation99,103,104 It appears that the adipogenic differentiation of SV cells is the fastest in mice, followed by swine and the slowest in cattle.

Adipocyte Development within IM

The rate of accretion and mode of cellular development of meat animal adipose tissue is depot and species dependent.Citation105 In contrast to other depots, IM adipocyte hyperplasia in meat animals apparently continues on with either a delayed plateau or no plateau in hyperplasia.Citation106 Intramuscular adipocyte size and number dictate skeletal muscle lipid content in several meat animals, including cattle, rabbits, and pigs.Citation107 Long-term selection for decreased backfat in pigs was highly associated with decreased IM as well.Citation108 Selection for lean growth in sheep also significantly reduced IM.Citation109 A study of cattle selected for high growth indicated no change in IM lipid despite a significant increase in backfat accretion.Citation110 Therefore, changes in growth with increased leanness may not be involved in the link between leanness and IM lipid. Intramuscular fat can be influenced by many factors () including age, diet, gender, fasted state, genetics, and physical activity.Citation113,128 These factors influence IM lipid accretion in a muscle- and species-dependent manner.Citation111 For example, somatotropin treatment in pigs produced depot dependent changes in fatty acid composition in a comparison of SQF and IM adipose tissue.Citation129 Dietary induced obesity increases IM lipid content in several species.Citation130 There is considerable evidence that diet/dietary components can influence IM lipid in production animals (). Reduced protein diets may increase IM lipid by preferentially increasing steroyl-CoA desaturase enzyme activity.Citation131 Muscle lipid content is increased by 40% when pigs are fed low lysine diets (). Changes in muscle lipid content could be due to age or altered muscle metabolism.

Table 2. Intramuscular lipid (IML) accretion that occurs naturally is species and muscle dependent

Generally, leptin gene expression marks SQF and distinguishes adipose depots from one another in several species.Citation105 Throughout growth, IM adipocytes are less metabolically activity when compared with SQF and perirenal adipocytes.Citation105 Lipogenesis distinguished SQF from IM adipose tissue explants and SV cell cultures from several breeds of cattle.Citation132 Gene expression for enzymes follows this pattern suggesting marked metabolic differences in adipocytes from different depots may exist at the gene level.Citation107 There may be a developmental delay in metabolic efficiency, hyperplasic or hypertrophic growth which contributes to these metabolic differences between IM and SQF.Citation105

Intramyocellular Lipid Dynamics

The idea that skeletal muscle, itself, may adsorb lipid and form intramyocellular lipid droplets is not new, but rarely constitutes a major source of lipid storage/available energy. Moreover, slow-oxidative skeletal muscle is the myofiber type that possesses sufficient metabolic and biochemical machinery (mitochondria) to process internal lipid droplets. As such, dramatic rises in intramyocellular lipid is commonly associated with pathological conditions such as insulin resistance.Citation78 Due to its presence in obesity, metabolic syndrome and other metabolic pathologies, intramyocellular lipid accumulation is beginning to be looked at with some interest. Few reports exist whereby this phenomenon even is evident in domesticated (meat) animals, but three studies are available evaluating this event in pigs.Citation133-135 In a study utilizing pigs of different body composition, Reiter et al. demonstrated that genetic preference for leanness resulted in lower genetic markers for variables of adiposity and lipid metabolism within adipose tissue over pigs of more conventional body composition.Citation134 However, select skeletal muscle markers for oxidative metabolism of lipids were higher in the leaner pig group; suggesting that intramyocellular lipid is processed more rapidly in lean-type pigs over conventional pigs of similar age and such metabolism is heightened by application of commercially available β-adrenergic agents.Citation134 Utilizing genetically small pigs, which are isolated from other breeds, it was shown that diets to mimic dyslipidemic metabolic syndrome resulted in greater accumulation of intramyocellular lipid, but did not result in greater overall numbers of lipid droplets in pigs.Citation133 Moreover, skeletal muscle contractile proteins appear altered when excess lipid was incorporated within afflicted skeletal muscle fibers.Citation133 At a more molecular level, Gandolfi et al. demonstrated that PLIN protein 2, was associated with lipid droplets in skeletal muscles of pigs displaying high IM lipid content, and that the expression of PLIN was associated with extra-cellular lipid availability to skeletal muscle cells.Citation135

Gene Expression Associations/Interactions in the Deposition Process of IM in Animals

Elucidation of common adipogenic mechanisms regulating depot-specific partitioning of lipids in animals and humans is important. In domestic animals, deposition of IM is regulated by complex interactions involving muscle, fat, and connective tissue and is associated with the genetic background, nutrition and development of an animal, as described in previous excellent reviews.Citation41,111,Citation136-140 Comprehensive high throughput genomic technologies have revealed large numbers of differentially expressed (DE) genes, and signaling pathways including adipogenic and lipogenic related genes, metabolic enzymes, and cholesterol and bile acid homeostasis () related to IM in livestock species.Citation111,Citation158-161 The transcriptome profiling of marbling LM tissue of heifers from Wagyu × Hereford indicated a strong positive correlation between expression of several adipogenic genes, adiponectin, C1Q, and collagen domain containing (ADIPOQ), SCD, thyroid hormone responsive (THRSP), and Fatty acid synthase (FAS), with IM.Citation161

Table 3. Common genes/genetic markers associated with INTMF and IM fat in humans and animals

Global analysis with GeneChip Bovine Genome array in Japanese Black and Holstein steers revealed that vast majority of DE genes were downregulated in IM and the mRNA abundance of adipogenic key regulators peroxisome proliferator-activated receptor gamma (PPARγ) and fatty acid binding protein (FABP)4, was neither directly associated with the size of adipocytes in a tissue nor with the amount of IM content.Citation162 On the contrary, higher expression of Zfp423, PPARγ, C/EBPα, and FABP4 in Wagyu with high marbling was observed as compared with the Angus cattle, indicating enhanced proliferation and/or adipogenic differentiation. Recently, Zfp423 was demonstrated as a very early marker for adipogenesis, which induces adipogenic commitment through upregulation of PPARγ expression.Citation51,52,97 The PPAR and mitogen-activated protein kinase (MAPK) signaling genes have also been positively correlated with the IM deposition in Beijing-you, a Chinese chicken breed possessing high IM as compared with the commercial broiler chicken Arbor Acres.Citation144,163

In pigs, positive association between the IM content and the expressions of myosin (MYL1), adipose-specific phospholipase A2 (AdPLA), melanocortin 4 receptor (MC4R), phosphoenolpyruvated carboxykinase (PEPCK), and SCD, and novel porcine FLJ36031 (pFLJ) genes are suggestive of their role in fast glycolysis and lipid deposition.Citation142,145 Genome-wide-association studies (GWAS) and the liver RNA-Seq on female pigs with extreme phenotypes for IM showed differential expression of transposable elements, long non-coding RNAs.Citation151,164,165 Furthermore, these studies established the role of putative protein-coding genes including Endotoxin lipopolysaccharide/pro-inflammatory cytokines (LPS/IL-1) in mediating inhibition of retinoid X receptors (RXR) function to regulate cholesterol homeostasis and bile acid homeostasis (). Besides changes in abundance of markers for lipid synthesis and transport, alterations have also been observed in protein and amino acid metabolism with differences in IM content.Citation111,166,167 Insulin growth factor-1 (IGF-1) and IGFBP3 have been associated with increased muscle growth of cattle resulting in elevation in metabolic enzymes with greater IM, suggesting an increased nutrient utilization through amino acid metabolism.Citation149

Genetic Variations/QTL and Epigenetic Changes Associated with IM Deposition in Animals

Several quantitative trait loci (QTL) regions and genetic markers for IM have been identified in pigs, chickens and cattle.Citation111,136 Some of excellent indicators for predicting IM content in livestock species include the SNPs in various genes e.g., DGAT1, DGAT2, serine-arginine-rich protein (SFRS18), adipogenic genes (ADRP, PPARγ), lipogenic genes (FASN, SREBP1, and SCD1), FABP4, PNAS-4, and calpain-system.Citation111,Citation168-170 QTL analysis has identified the region between S0228 and SW1881on pig chromosome 6 (SSC6) that influences IM.Citation171,172

Recent studies have suggested that the effect of polymorphisms influencing IM, cholesterol, and fatty acid contents are modulated by several factors related to muscle location, metabolism and function.Citation173 Stearoyl-CoA desaturase (SCD1) activity has been shown to increase with higher fat accumulation in skeletal muscle in humans and animals possibly related to the SNPs in 3¢-UTR that contains G > A and C > T in the bovine gene (which are also highly conserved in humans and pigs).Citation141 Furthermore, the polymorphism in the PPARγ (TA haplotype favorable over CG haplotype) upstream transcriptional regulatory region has been associated with higher IM.Citation174 On the contrary, a few other GWA have had limited success in the identification of the genetic drivers of IM percentage which could be due to variations including treatments, genetics etc., resulting in the similar phenotype.Citation175,176 High throughput genomic technologies using powerful SNP panels will have to be combined with phenomics approach to further understand the complexities of IM.

During the past decade, various miRNAs, endogenous small non-coding RNAs, have been shown to impact on epigenetic regulatory mechanisms with implications to the global gene expression in adipogenesis in cattle, pigs, and lipid metabolism disorders.Citation101,Citation177-183 In pigs, eight inflammation-related and nine diabetes-related miRNAs were present in higher abundance in the INTMF transcriptome compared with the SQF indicating the metabolic role of IM in obesity-related metabolic dysfunction.Citation75 The expansion of INTMF could induce some changes in muscle metabolism resulting in insulin sensitivity because of the release of adipokines and metabolites from fat cells surrounding muscle fibers.Citation77 Future studies focusing on the identification of the factors that modulate miRNAs expression could further expand our current understanding of the environment and genetic factors that influence gene expression during adipogenesis.

Common Genes/Genetic Markers Associated with IM and IM Fat in Humans and Animals

Emerging evidence suggests that INTMF accumulation surrounding skeletal muscle in humans has significant negative impacts on health, causing obesity/type 2 diabetes and their associated conditions.Citation138,184 To understand the genetic complexity of these conditions, many studies have focused on human subjects or on the mouse as a model organism.Citation77 Accumulation of INTMF fatty acid metabolites in human skeletal muscle plays an important role in the regulation of pyruvate dehydrogenase kinase (PDK4) gene expression possibly through interaction with PPARγ and PPARγ coactivator 1α (PGC1α).Citation143,185 Association of human minor alleles for nonsynonymous coding variants (G531L, I66V) in the carnitine palmitoyltransferase-1B gene (CPT1B, the rate limiting enzyme in mitochondrial β-oxidation of long-chain fatty acids), has been associated with lower IM and higher SQ, independent of overall adiposity.Citation82

Several studies have established the role of human agouti signaling protein (ASIP) expressed in adipose tissue, in the development of metabolic disorders like obesity and insulin resistance through influencing the expression of STAT 1/3 (signal transducer and activator of transcription) and PPARγ transcription factors.Citation155 ASIP/melanocortin signaling system has also been found to regulate lipid metabolism in chickens and ASIP mRNA abundance showed more than 9-fold increase in IM of Japanese Black cattle compared with Holstein ().Citation156,157 Three independent whole genome association studies in humans provided evidence that three different SNPs in the FTO (fat mass and obesity-associated) gene intron 1 are significantly associated with both childhood and adult obesity.Citation153,186,187 FTO gene participates in fat tissue and energy homeostasis and adults homozygous for the risk alleles of SNP rs993939 are linked with increased risk for obesity.Citation153 Interestingly, the FTO gene was assigned between SW1302 and SWR1130 on SSC6 in the porcine (), adjacent to the region with a number of QTLs for average daily gain, IM percentage and body lipid content.Citation154 Comprehensive database search revealed that FLJ, which are involved in fat deposition and adipogenesis are highly conserved among different species and pig FLJ (pFLJ) shares 93%, 83%, 92%, and 92% homology with human, mouse, chimpanzee and rhesus monkey FLJ, respectively.Citation145 FLJ20920 is an important gene in human adipogenesis through its interaction with PPARγ.Citation146 In pigs, the Lipin1 gene may have a crucial effect on body lipid accumulation, whereas the Lipin-β isoform may play an important role in IM deposition in obese pigs.Citation147 Overexpression of PLIN2 has been shown to result in decreased gene expression of several PPARα target genes and reduced transcriptional activity of mitochondrial genes facilitating intramyocellular lipid (IMCL) storage.Citation148 A better understanding of the molecular pathways or genetic markers involved in INTMF and IM deposition could potentially result in novel therapies to treat obesity.

Conclusions

As summarized in , review of human INTMF studies indicated that INTMF is influenced by aging, physical activity, race and obesity and induced limb inactivity, and is increased with weight loss, weight gain, or with no weight change. As such, INTMF is clearly a threat to type 2 diabetes and projects many physiologies similar to VAT. Studies of INTMF in animals are not numerous and covered different topics such as quantitative deposition and genetic relationships with other fat depots. Therefore, in growing animals, INTMF is regulated by genetic means distinct from the other fat depots. The relationship between leanness in pigs and higher proportions of INTMF fat was not similarly observed in humans. More specifically, in humans, changes in muscle mass, strength, and quality with aging are associated with INTMF accretion whereas muscle characteristics do not change with age pigs. It seems as though pigs are not “old” relative to humans. Even so, gene expression profiling and intrinsic methylation differences in pigs demonstrated that INTMF and VAT are primarily associated with inflammatory and immune processes. In humans INTMF is closely linked to VAT and INTMF was a good predictor for insulin sensitivity nearly equal to that of VAT. Therefore, it seems that in the pig and humans, INTMF and VAT share a similar pattern in distribution and association of components dictating insulin sensitivity.

Figure 1. Major metabolic, physiological, and genetic changes associated with different Intermuscular and Intramuscular fat in animals and humans.

Figure 1. Major metabolic, physiological, and genetic changes associated with different Intermuscular and Intramuscular fat in animals and humans.

Intramuscular adipocyte development affects the quality of meat, and studies in meat animals have included the influence of aging, gender, diet, obesity, and birth weights. Gene expression analysis and genetic analysis have identified candidate genes involved in IM adipocyte development in these animals. Alternatively, IM adipocyte development in muscle of humans is rare, and rarely seen except during some pathological circumstance, such as lipotrophic muscular atrophy. Therefore the cellularity and regulation of human IM are rarely studied.

Several genetic links between human and meat animal adipogenesis have been identified. The human agouti signaling protein (ASIP) is expressed in adipose tissue () and may be involved in the development of obesity and insulin resistance by influencing the expression of STAT 1/3 and PPARγ transcription factors.Citation155 ASIP mRNA was increased in IM more than 9-fold in Japanese Black cattle compared with Holstein.Citation157 Whole genome association studies in humans provided evidence that three different SNPs in the FTO gene are associated with both childhood and adult obesity.Citation153,186,187 The FTO gene is located on porcine SSC6 (), adjacent to a number of QTLs for IM percentage and body lipid content.Citation154

Comprehensive database search revealed that FLJ, which are involved in fat deposition and adipogenesis are highly conserved among different species including the pig FLJ (pFL).Citation145 FLJ20920 is involved in human adipogenesis through its interaction PPARγ.Citation146 In pigs, the PLIN1 and PLIN2 genes have strong genetic effects on fat depot lipid accretion including IM accumulation, and the PLIN-β isoform may influence IM deposition in obese pigs ().Citation147 PLIN1 deficiency results in immature adipocyte development human lipodystrophy.Citation188 In humans, overexpression of PLIN 2 decreases gene expression of several PPARα target genes and facilitates intramyocellular lipid.Citation148 In pigs, PLIN2 gene expression was associated with IM deposition.Citation72,135 It is clear that the PLIN family may influence intramuscular lipid regardless of species.

The literature clearly indicated that there is very little IMCL in meat animal muscle and very little evidence of intramuscular adipocytes in human muscle. In meat animals IMCL is present in newborn and young animals but dissipates as intramuscular adipocytes develop and increase in number.Citation106 Possibly, intramuscular adipocytes prevent or antagonize IMCL in animals to prevent lipotoxicity in muscle.Citation130 The virtual absence of intramuscular adipocytes in human muscle may augment IMCL and associated dire consequences.Citation130 Future studies are dictated to validate these possibilities.

Disclosure of Potential Conflicts of Interest

No potential conflicts of interest were disclosed.

References

  • Rosenquist KJ, Pedley A, Massaro JM, Therkelsen KE, Murabito JM, Hoffmann U, Fox CS. Visceral and subcutaneous fat quality and cardiometabolic risk. JACC Cardiovasc Imaging 2013; 6:762-71; PMID:23664720; http://dx.doi.org/10.1016/j.jcmg.2012.11.021
  • Matsuzawa Y. The role of fat topology in the risk of disease. Int J Obes (Lond) 2008; 32(Suppl 7):S83-92; PMID:19136997; http://dx.doi.org/10.1038/ijo.2008.243
  • Matsuzawa Y. White adipose tissue and cardiovascular disease. Best Pract Res Clin Endocrinol Metab 2005; 19:637-47; PMID:16311222; http://dx.doi.org/10.1016/j.beem.2005.07.001
  • Hocking S, Samocha-Bonet D, Milner KL, Greenfield JR, Chisholm DJ. Adiposity and insulin resistance in humans: the role of the different tissue and cellular lipid depots. Endocr Rev 2013; 34:463-500; PMID:23550081; http://dx.doi.org/10.1210/er.2012-1041
  • Alvehus M, Burén J, Sjöström M, Goedecke J, Olsson T. The human visceral fat depot has a unique inflammatory profile. Obesity (Silver Spring) 2010; 18:879-83; PMID:20186138; http://dx.doi.org/10.1038/oby.2010.22
  • Matsuzawa Y. Adiponectin: Identification, physiology and clinical relevance in metabolic and vascular disease. Atheroscler Suppl 2005; 6:7-14; PMID:15823491; http://dx.doi.org/10.1016/j.atherosclerosissup.2005.02.003
  • Zhang Y, Zitsman JL, Hou J, Fennoy I, Guo K, Feinberg J, Leibel RL. Fat cell size and adipokine expression in relation to gender, depot, and metabolic risk factors in morbidly obese adolescents. Obesity (Silver Spring) 2014; 22:691-7; PMID:23804589; http://dx.doi.org/10.1002/oby.20528
  • Patel P, Abate N. Body fat distribution and insulin resistance. Nutrients 2013; 5:2019-27; PMID:23739143; http://dx.doi.org/10.3390/nu5062019
  • Wueest S, Schoenle EJ, Konrad D. Depot-specific differences in adipocyte insulin sensitivity in mice are diet- and function-dependent. Adipocyte 2012; 1:153-6; PMID:23700524; http://dx.doi.org/10.4161/adip.19910
  • Joffe YT, Collins M, Goedecke JH. The relationship between dietary fatty acids and inflammatory genes on the obese phenotype and serum lipids. Nutrients 2013; 5:1672-705; PMID:23698162; http://dx.doi.org/10.3390/nu5051672
  • Goedecke JH, Levitt NS, Evans J, Ellman N, Hume DJ, Kotze L, Tootla M, Victor H, Keswell D. The role of adipose tissue in insulin resistance in women of African ancestry. J Obes 2013; 2013:952916; PMID:23401754; http://dx.doi.org/10.1155/2013/952916
  • Billon N, Monteiro MC, Dani C. Developmental origin of adipocytes: new insights into a pending question. Biol Cell 2008; 100:563-75; PMID:18793119; http://dx.doi.org/10.1042/BC20080011
  • Billon N, Dani C. Developmental origins of the adipocyte lineage: new insights from genetics and genomics studies. Stem Cell Rev 2012; 8:55-66; PMID:21365256; http://dx.doi.org/10.1007/s12015-011-9242-x
  • Sanchez-Gurmaches J, Guertin DA. Adipocyte lineages: Tracing back the origins of fat. Biochim Biophys Acta 2014;1842:340-51; PMID:23747579; http://dx.doi.org/10.1016/j.bbadis.2013.05.027
  • McLaughlin T, Lamendola C, Liu A, Abbasi F. Preferential fat deposition in subcutaneous versus visceral depots is associated with insulin sensitivity. J Clin Endocrinol Metab 2011; 96:E1756-60; PMID:21865361; http://dx.doi.org/10.1210/jc.2011-0615
  • Bidulescu A, Liu J, Hickson DA, Hairston KG, Fox ER, Arnett DK, Sumner AE, Taylor HA, Gibbons GH. Gender differences in the association of visceral and subcutaneous adiposity with adiponectin in African Americans: the Jackson Heart Study. BMC Cardiovasc Disord 2013; 13:9; PMID:23433085; http://dx.doi.org/10.1186/1471-2261-13-9
  • Misra A, Garg A, Abate N, Peshock RM, Stray-Gundersen J, Grundy SM. Relationship of anterior and posterior subcutaneous abdominal fat to insulin sensitivity in nondiabetic men. Obes Res 1997; 5:93-9; PMID:9112243; http://dx.doi.org/10.1002/j.1550-8528.1997.tb00648.x
  • Snijder MB, Dekker JM, Visser M, Bouter LM, Stehouwer CD, Kostense PJ, Yudkin JS, Heine RJ, Nijpels G, Seidell JC. Associations of hip and thigh circumferences independent of waist circumference with the incidence of type 2 diabetes: the Hoorn Study. Am J Clin Nutr 2003; 77:1192-7; PMID:12716671
  • Tran TT, Yamamoto Y, Gesta S, Kahn CR. Beneficial effects of subcutaneous fat transplantation on metabolism. Cell Metab 2008; 7:410-20; PMID:18460332; http://dx.doi.org/10.1016/j.cmet.2008.04.004
  • Preis SR, Massaro JM, Robins SJ, Hoffmann U, Vasan RS, Irlbeck T, Meigs JB, Sutherland P, D’Agostino RB Sr., O’Donnell CJ, et al. Abdominal subcutaneous and visceral adipose tissue and insulin resistance in the Framingham heart study. Obesity (Silver Spring) 2010; 18:2191-8; PMID:20339361; http://dx.doi.org/10.1038/oby.2010.59
  • Kishida K, Funahashi T, Matsuzawa Y, Shimomura I. Visceral adiposity as a target for the management of the metabolic syndrome. Ann Med 2012; 44:233-41; PMID:21612331; http://dx.doi.org/10.3109/07853890.2011.564202
  • Matsuzawa Y, Funahashi T, Nakamura T. The concept of metabolic syndrome: contribution of visceral fat accumulation and its molecular mechanism. J Atheroscler Thromb 2011; 18:629-39; PMID:21737960; http://dx.doi.org/10.5551/jat.7922
  • Matsuzawa Y, Shimomura I, Nakamura T, Keno Y, Kotani K, Tokunaga K. Pathophysiology and pathogenesis of visceral fat obesity. Obes Res 1995; 3(Suppl 2):187S-94S; PMID:8581775; http://dx.doi.org/10.1002/j.1550-8528.1995.tb00462.x
  • Shay CM, Carnethon MR, Church TR, Hankinson AL, Chan C, Jacobs DR Jr., Lewis CE, Schreiner PJ, Sternfeld B, Sidney S. Lower extremity fat mass is associated with insulin resistance in overweight and obese individuals: the CARDIA study. Obesity (Silver Spring) 2011; 19:2248-53; PMID:21617639; http://dx.doi.org/10.1038/oby.2011.113
  • Evans J, Goedecke JH, Söderström I, Burén J, Alvehus M, Blomquist C, Jonsson F, Hayes PM, Adams K, Dave JA, et al. Depot- and ethnic-specific differences in the relationship between adipose tissue inflammation and insulin sensitivity. Clin Endocrinol (Oxf) 2011; 74:51-9; PMID:20874774; http://dx.doi.org/10.1111/j.1365-2265.2010.03883.x
  • Manolopoulos KN, Karpe F, Frayn KN. Gluteofemoral body fat as a determinant of metabolic health. Int J Obes (Lond) 2010; 34:949-59; PMID:20065965; http://dx.doi.org/10.1038/ijo.2009.286
  • Lovejoy JC, Smith SR, Rood JC. Comparison of regional fat distribution and health risk factors in middle-aged white and African American women: The Healthy Transitions Study. Obes Res 2001; 9:10-6; PMID:11346662; http://dx.doi.org/10.1038/oby.2001.2
  • Lovejoy JC, de la Bretonne JA, Klemperer M, Tulley R. Abdominal fat distribution and metabolic risk factors: effects of race. Metabolism 1996; 45:1119-24; PMID:8781299; http://dx.doi.org/10.1016/S0026-0495(96)90011-6
  • Goedecke JH, Evans J, Keswell D, Stimson RH, Livingstone DE, Hayes P, Adams K, Dave JA, Victor H, Levitt NS, et al. Reduced gluteal expression of adipogenic and lipogenic genes in Black South African women is associated with obesity-related insulin resistance. J Clin Endocrinol Metab 2011; 96:E2029-33; PMID:21956425; http://dx.doi.org/10.1210/jc.2011-1576
  • Goedecke JH, Levitt NS, Lambert EV, Utzschneider KM, Faulenbach MV, Dave JA, West S, Victor H, Evans J, Olsson T, et al. Differential effects of abdominal adipose tissue distribution on insulin sensitivity in black and white South African women. Obesity (Silver Spring) 2009; 17:1506-12; PMID:19300428; http://dx.doi.org/10.1038/oby.2009.73
  • Tittelbach TJ, Berman DM, Nicklas BJ, Ryan AS, Goldberg AP. Racial differences in adipocyte size and relationship to the metabolic syndrome in obese women. Obes Res 2004; 12:990-8; PMID:15229339; http://dx.doi.org/10.1038/oby.2004.121
  • Staiano AE, Broyles ST, Gupta AK, Katzmarzyk PT. Ethnic and sex differences in visceral, subcutaneous, and total body fat in children and adolescents. Obesity (Silver Spring) 2013;21:1251-5; PMID:23670982; http://dx.doi.org/10.1002/oby.20210
  • Smith SR, Lovejoy JC, Greenway F, Ryan D, deJonge L, de la Bretonne J, Volafova J, Bray GA. Contributions of total body fat, abdominal subcutaneous adipose tissue compartments, and visceral adipose tissue to the metabolic complications of obesity. Metabolism 2001; 50:425-35; PMID:11288037; http://dx.doi.org/10.1053/meta.2001.21693
  • White UA, Tchoukalova YD. Sex dimorphism and depot differences in adipose tissue function. Biochim Biophys Acta 2014;1842:377-9; PMID:23684841; http://dx.doi.org/10.1016/j.bbadis.2013.05.006
  • da Costa AS, Pires VM, Fontes CM, Mestre Prates JA. Expression of genes controlling fat deposition in two genetically diverse beef cattle breeds fed high or low silage diets. BMC Vet Res 2013; 9:118; PMID:23767408; http://dx.doi.org/10.1186/1746-6148-9-118
  • Vasconcelos JT, Sawyer JE, Tedeschi LO, McCollum FT, Greene LW. Effects of different growing diets on performance, carcass characteristics, insulin sensitivity, and accretion of intramuscular and subcutaneous adipose tissue of feedlot cattle. J Anim Sci 2009; 87:1540-7; PMID:19098228; http://dx.doi.org/10.2527/jas.2008-0934
  • Reuter RR, Beck PA. Southern Section Interdisciplinary Beef Cattle Symposium: Carryover effects of stocker cattle systems on feedlot performance and carcass characteristics. J Anim Sci 2013; 91:508-15; PMID:23048147; http://dx.doi.org/10.2527/jas.2012-5527
  • Mello AS Jr., Jenschke BE, Senaratne LS, Carr TP, Erickson GE, Calkins CR. Effects of feeding modified distillers grains plus solubles on marbling attributes, proximate composition, and fatty acid profile of beef. J Anim Sci 2012; 90:4634-40; PMID:22859752; http://dx.doi.org/10.2527/jas.2010-3240
  • Schmidt JR, Miller MC, Andrae JG, Ellis SE, Duckett SK. Effect of summer forage species grazed during finishing on animal performance, carcass quality, and meat quality. J Anim Sci 2013; 91:4451-61; PMID:23825343; http://dx.doi.org/10.2527/jas.2012-5405
  • Sharman ED, Lancaster PA, Krehbiel CR, Hilton GG, Stein DR, Desilva U, Horn GW. Effects of starch- vs. fiber-based energy supplements during winter grazing on partitioning of fat among depots and adipose tissue gene expression in growing cattle and final carcass characteristics. J Anim Sci 2013; 91:2264-77; PMID:23463572; http://dx.doi.org/10.2527/jas.2012-5284
  • Du M, Yin J, Zhu MJ. Cellular signaling pathways regulating the initial stage of adipogenesis and marbling of skeletal muscle. Meat Sci 2010; 86:103-9; PMID:20510530; http://dx.doi.org/10.1016/j.meatsci.2010.04.027
  • Stern HM, Brown AM, Hauschka SD. Myogenesis in paraxial mesoderm: preferential induction by dorsal neural tube and by cells expressing Wnt-1. Development 1995; 121:3675-86; PMID:8582280
  • Cossu G, Borello U. Wnt signaling and the activation of myogenesis in mammals. EMBO J 1999; 18:6867-72; PMID:10601008; http://dx.doi.org/10.1093/emboj/18.24.6867
  • Du M, Zhao JX, Yan X, Huang Y, Nicodemus LV, Yue W, McCormick RJ, Zhu MJ. Fetal muscle development, mesenchymal multipotent cell differentiation, and associated signaling pathways. J Anim Sci 2011; 89:583-90; PMID:20852073; http://dx.doi.org/10.2527/jas.2010-3386
  • Atit R, Sgaier SK, Mohamed OA, Taketo MM, Dufort D, Joyner AL, Niswander L, Conlon RA. Beta-catenin activation is necessary and sufficient to specify the dorsal dermal fate in the mouse. Dev Biol 2006; 296:164-76; PMID:16730693; http://dx.doi.org/10.1016/j.ydbio.2006.04.449
  • Kajimura S, Seale P, Kubota K, Lunsford E, Frangioni JV, Gygi SP, Spiegelman BM. Initiation of myoblast to brown fat switch by a PRDM16-C/EBP-beta transcriptional complex. Nature 2009; 460:1154-8; PMID:19641492; http://dx.doi.org/10.1038/nature08262
  • Tseng YH, Kokkotou E, Schulz TJ, Huang TL, Winnay JN, Taniguchi CM, Tran TT, Suzuki R, Espinoza DO, Yamamoto Y, et al. New role of bone morphogenetic protein 7 in brown adipogenesis and energy expenditure. Nature 2008; 454:1000-4; PMID:18719589; http://dx.doi.org/10.1038/nature07221
  • Taga H, Chilliard Y, Picard B, Zingaretti MC, Bonnet M. Foetal bovine intermuscular adipose tissue exhibits histological and metabolic features of brown and white adipocytes during the last third of pregnancy. Animal 2012; 6:641-9; PMID:22436281; http://dx.doi.org/10.1017/S1751731111001716
  • Taga H, Bonnet M, Picard B, Zingaretti MC, Cassar-Malek I, Cinti S, Chilliard Y. Adipocyte metabolism and cellularity are related to differences in adipose tissue maturity between Holstein and Charolais or Blond d’Aquitaine fetuses. J Anim Sci 2011; 89:711-21; PMID:21036936; http://dx.doi.org/10.2527/jas.2010-3234
  • MacDougald OA, Mandrup S. Adipogenesis: forces that tip the scales. Trends Endocrinol Metab 2002; 13:5-11; PMID:11750856; http://dx.doi.org/10.1016/S1043-2760(01)00517-3
  • Gupta RK, Arany Z, Seale P, Mepani RJ, Ye L, Conroe HM, Roby YA, Kulaga H, Reed RR, Spiegelman BM. Transcriptional control of preadipocyte determination by Zfp423. Nature 2010; 464:619-23; PMID:20200519; http://dx.doi.org/10.1038/nature08816
  • Gupta RK, Mepani RJ, Kleiner S, Lo JC, Khandekar MJ, Cohen P, Frontini A, Bhowmick DC, Ye L, Cinti S, et al. Zfp423 expression identifies committed preadipocytes and localizes to adipose endothelial and perivascular cells. Cell Metab 2012; 15:230-9; PMID:22326224; http://dx.doi.org/10.1016/j.cmet.2012.01.010
  • Huang Y, Das AK, Yang QY, Zhu MJ, Du M. Zfp423 promotes adipogenic differentiation of bovine stromal vascular cells. PLoS One 2012; 7:e47496; PMID:23071815; http://dx.doi.org/10.1371/journal.pone.0047496
  • Avram MM, Avram AS, James WD. Subcutaneous fat in normal and diseased states 3. Adipogenesis: from stem cell to fat cell. J Am Acad Dermatol 2007; 56:472-92; PMID:17317490; http://dx.doi.org/10.1016/j.jaad.2006.06.022
  • Fajas L, Debril MB, Auwerx J. Peroxisome proliferator-activated receptor-gamma: from adipogenesis to carcinogenesis. J Mol Endocrinol 2001; 27:1-9; PMID:11463572; http://dx.doi.org/10.1677/jme.0.0270001
  • Spiegelman BM, Flier JS. Adipogenesis and obesity: rounding out the big picture. Cell 1996; 87:377-89; PMID:8898192; http://dx.doi.org/10.1016/S0092-8674(00)81359-8
  • Rosen ED, MacDougald OA. Adipocyte differentiation from the inside out. Nat Rev Mol Cell Biol 2006; 7:885-96; PMID:17139329; http://dx.doi.org/10.1038/nrm2066
  • Brun RP, Spiegelman BM. PPAR gamma and the molecular control of adipogenesis. J Endocrinol 1997; 155:217-8; PMID:9415052; http://dx.doi.org/10.1677/joe.0.1550217
  • Allen CE, Beitz DC, Cramer DA, Kauffman RG. Biology of fat in meat animals. Madison: University of Wisconsin-Madison1976 Contract No.: No. 234.
  • Haugebak CD, Hedrick HB, Asplund JM. Relationship between extramuscular adipose tissue lipoprotein lipase activity and intramuscular lipid deposition in fattening lambs. J Anim Sci 1974; 39:1026-31; PMID:4443312
  • Robelin J. Growth of adipose tissues in cattle; partitioning between depots, chemical composition and cellularity. A review. Livest Prod Sci 1986; 14:349-64; http://dx.doi.org/10.1016/0301-6226(86)90014-X
  • Kouba M, Sellier P. A review of the factors influencing the development of intermuscular adipose tissue in the growing pig. Meat Sci 2011; 88:213-20; PMID:21303725; http://dx.doi.org/10.1016/j.meatsci.2011.01.003
  • Cianzio DS, Topel DG, Whitehurst GB, Beitz DC, Self HL. Adipose tissue growth and cellularity: changes in bovine adipocyte size and number. J Anim Sci 1985; 60:970-6; PMID:3988658
  • Kouba M, Bonneau M, Noblet J. Relative development of subcutaneous, intermuscular, and kidney fat in growing pigs with different body compositions. J Anim Sci 1999; 77:622-9; PMID:10229357
  • Gispert M, Font I Furnols M, Gil M, Velarde A, Diestre A, Carrión D, Sosnicki AA, Plastow GS. Relationships between carcass quality parameters and genetic types. Meat Sci 2007; 77:397-404; PMID:22061793; http://dx.doi.org/10.1016/j.meatsci.2007.04.006
  • Eggert JM, Grant AL, Schinckel AP. Factors Affecting Fat Distribution in Pork Carcasses. Prof Anim Sci 2007; 23:42-53
  • Wood JD, Richardson RI, Nute GR, Fisher AV, Campo MM, Kasapidou E, Sheard PR, Enser M. Effects of fatty acids on meat quality: a review. Meat Sci 2004; 66:21-32; PMID:22063928; http://dx.doi.org/10.1016/S0309-1740(03)00022-6
  • Hoque MA, Suzuki K, Kadowaki H, Shibata T, Oikawa T. Genetic parameters for feed efficiency traits and their relationships with growth and carcass traits in Duroc pigs. J Anim Breed Genet 2007; 124:108-16; PMID:17550351; http://dx.doi.org/10.1111/j.1439-0388.2007.00650.x
  • Hoque MA, Katoh K, Suzuki K. Genetic associations of residual feed intake with serum insulin-like growth factor-I and leptin concentrations, meat quality, and carcass cross sectional fat area ratios in Duroc pigs. J Anim Sci 2009; 87:3069-75; PMID:19465494; http://dx.doi.org/10.2527/jas.2008-1268
  • Suzuki K, Inomata K, Katoh K, Kadowaki H, Shibata T. Genetic correlations among carcass cross-sectional fat area ratios, production traits, intramuscular fat, and serum leptin concentration in Duroc pigs. J Anim Sci 2009; 87:2209-15; PMID:19329483; http://dx.doi.org/10.2527/jas.2008-0866
  • Bergen R, Miller SP, Wilton JW, Mandell IB. Genetic correlations between live yearling bull and steer carcass traits adjusted to different slaughter end points. 2. Carcass fat partitioning. J Anim Sci 2006; 84:558-66; PMID:16478947
  • Davoli R, Gandolfi G, Braglia S, Comella M, Zambonelli P, Buttazzoni L, Russo V. New SNP of the porcine perilipin 2 (PLIN2) gene, association with carcass traits and expression analysis in skeletal muscle. Mol Biol Rep 2011; 38:1575-83; PMID:20842447; http://dx.doi.org/10.1007/s11033-010-0266-0
  • Zhou C, Zhang J, Ma J, Jiang A, Tang G, Mai M, Zhu L, Bai L, Li M, Li X. Gene expression profiling reveals distinct features of various porcine adipose tissues. Lipids Health Dis 2013; 12:75; PMID:23705929; http://dx.doi.org/10.1186/1476-511X-12-75
  • Li M, Wu H, Wang T, Xia Y, Jin L, Jiang A, Zhu L, Chen L, Li R, Li X. Co-methylated genes in different adipose depots of pig are associated with metabolic, inflammatory and immune processes. Int J Biol Sci 2012; 8:831-7; PMID:22719223; http://dx.doi.org/10.7150/ijbs.4493
  • Ma J, Yu S, Wang F, Bai L, Xiao J, Jiang Y, Chen L, Wang J, Jiang A, Li M, et al. MicroRNA Transcriptomes Relate Intermuscular Adipose Tissue to Metabolic Risk. Int J Mol Sci 2013; 14:8611-24; PMID:23609494; http://dx.doi.org/10.3390/ijms14048611
  • Cornier MA, Dabelea D, Hernandez TL, Lindstrom RC, Steig AJ, Stob NR, Van Pelt RE, Wang H, Eckel RH. The metabolic syndrome. Endocr Rev 2008; 29:777-822; PMID:18971485; http://dx.doi.org/10.1210/er.2008-0024
  • Vettor R, Milan G, Franzin C, Sanna M, De Coppi P, Rizzuto R, Federspil G. The origin of intermuscular adipose tissue and its pathophysiological implications. Am J Physiol Endocrinol Metab 2009; 297:E987-98; PMID:19738037; http://dx.doi.org/10.1152/ajpendo.00229.2009
  • Coen PM, Goodpaster BH. Role of intramyocelluar lipids in human health. Trends Endocrinol Metab 2012; 23:391-8; PMID:22721584; http://dx.doi.org/10.1016/j.tem.2012.05.009
  • Delmonico MJ, Harris TB, Visser M, Park SW, Conroy MB, Velasquez-Mieyer P, Boudreau R, Manini TM, Nevitt M, Newman AB, et al.; Health, Aging, and Body. Longitudinal study of muscle strength, quality, and adipose tissue infiltration. Am J Clin Nutr 2009; 90:1579-85; PMID:19864405; http://dx.doi.org/10.3945/ajcn.2009.28047
  • Manini TM, Clark BC, Nalls MA, Goodpaster BH, Ploutz-Snyder LL, Harris TB. Reduced physical activity increases intermuscular adipose tissue in healthy young adults. Am J Clin Nutr 2007; 85:377-84; PMID:17284732
  • Marcus RL, Addison O, Kidde JP, Dibble LE, Lastayo PC. Skeletal muscle fat infiltration: impact of age, inactivity, and exercise. J Nutr Health Aging 2010; 14:362-6; PMID:20424803; http://dx.doi.org/10.1007/s12603-010-0081-2
  • Miljkovic I, Yerges LM, Li H, Gordon CL, Goodpaster BH, Kuller LH, Nestlerode CS, Bunker CH, Patrick AL, Wheeler VW, et al. Association of the CPT1B gene with skeletal muscle fat infiltration in Afro-Caribbean men. Obesity (Silver Spring) 2009; 17:1396-401; PMID:19553926
  • Tuttle LJ, Sinacore DR, Mueller MJ. Intermuscular adipose tissue is muscle specific and associated with poor functional performance. J Aging Res 2012; 2012:172957; PMID:22666591; http://dx.doi.org/10.1155/2012/172957
  • Goodpaster BH, Chomentowski P, Ward BK, Rossi A, Glynn NW, Delmonico MJ, Kritchevsky SB, Pahor M, Newman AB. Effects of physical activity on strength and skeletal muscle fat infiltration in older adults: a randomized controlled trial. J Appl Physiol (1985) 2008; 105:1498-503; PMID:18818386; http://dx.doi.org/10.1152/japplphysiol.90425.2008
  • Miljkovic I, Cauley JA, Petit MA, Ensrud KE, Strotmeyer E, Sheu Y, Gordon CL, Goodpaster BH, Bunker CH, Patrick AL, et al.; Osteoporotic Fractures in Men Research Group; Tobago Health Studies Research Group. Greater adipose tissue infiltration in skeletal muscle among older men of African ancestry. J Clin Endocrinol Metab 2009; 94:2735-42; PMID:19454588; http://dx.doi.org/10.1210/jc.2008-2541
  • Li WZ, Zhao SM, Huang Y, Yang MH, Pan HB, Zhang X, Ge CR, Gao SZ. Expression of lipogenic genes during porcine intramuscular preadipocyte differentiation. Res Vet Sci 2012; 93:1190-4; PMID:22795880; http://dx.doi.org/10.1016/j.rvsc.2012.06.006
  • Boettcher M, Machann J, Stefan N, Thamer C, Häring HU, Claussen CD, Fritsche A, Schick F. Intermuscular adipose tissue (IMAT): association with other adipose tissue compartments and insulin sensitivity. J Magn Reson Imaging 2009; 29:1340-5; PMID:19422021; http://dx.doi.org/10.1002/jmri.21754
  • Du M, Huang Y, Das AK, Yang Q, Duarte MS, Dodson MV, Zhu MJ. Meat Science and Muscle Biology Symposium: manipulating mesenchymal progenitor cell differentiation to optimize performance and carcass value of beef cattle. J Anim Sci 2013; 91:1419-27; PMID:23100595; http://dx.doi.org/10.2527/jas.2012-5670
  • Uezumi A, Fukada S, Yamamoto N, Takeda S, Tsuchida K. Mesenchymal progenitors distinct from satellite cells contribute to ectopic fat cell formation in skeletal muscle. Nat Cell Biol 2010; 12:143-52; PMID:20081842; http://dx.doi.org/10.1038/ncb2014
  • Uezumi A, Ito T, Morikawa D, Shimizu N, Yoneda T, Segawa M, Yamaguchi M, Ogawa R, Matev MM, Miyagoe-Suzuki Y, et al. Fibrosis and adipogenesis originate from a common mesenchymal progenitor in skeletal muscle. J Cell Sci 2011; 124:3654-64; PMID:22045730; http://dx.doi.org/10.1242/jcs.086629
  • Joe AW, Yi L, Natarajan A, Le Grand F, So L, Wang J, Rudnicki MA, Rossi FM. Muscle injury activates resident fibro/adipogenic progenitors that facilitate myogenesis. Nat Cell Biol 2010; 12:153-63; PMID:20081841; http://dx.doi.org/10.1038/ncb2015
  • Zhu MJ, Ford SP, Means WJ, Hess BW, Nathanielsz PW, Du M. Maternal nutrient restriction affects properties of skeletal muscle in offspring. J Physiol 2006; 575:241-50; PMID:16763001; http://dx.doi.org/10.1113/jphysiol.2006.112110
  • Ma J, Yang J, Zhou L, Zhang Z, Ma H, Xie X, Zhang F, Xiong X, Cui L, Yang H, et al. Genome-wide association study of meat quality traits in a White Duroc×Erhualian F2 intercross and Chinese Sutai pigs. PLoS One 2013; 8:e64047; PMID:23724019; http://dx.doi.org/10.1371/journal.pone.0064047
  • Yan X, Huang Y, Zhao JX, Long NM, Uthlaut AB, Zhu MJ, Ford SP, Nathanielsz PW, Du M. Maternal obesity-impaired insulin signaling in sheep and induced lipid accumulation and fibrosis in skeletal muscle of offspring. Biol Reprod 2011; 85:172-8; PMID:21349823; http://dx.doi.org/10.1095/biolreprod.110.089649
  • Yan X, Zhu MJ, Xu W, Tong JF, Ford SP, Nathanielsz PW, Du M. Up-regulation of Toll-like receptor 4/nuclear factor-kappaB signaling is associated with enhanced adipogenesis and insulin resistance in fetal skeletal muscle of obese sheep at late gestation. Endocrinology 2010; 151:380-7; PMID:19887565; http://dx.doi.org/10.1210/en.2009-0849
  • Zhu MJ, Han B, Tong J, Ma C, Kimzey JM, Underwood KR, Xiao Y, Hess BW, Ford SP, Nathanielsz PW, et al. AMP-activated protein kinase signalling pathways are down regulated and skeletal muscle development impaired in fetuses of obese, over-nourished sheep. J Physiol 2008; 586:2651-64; PMID:18372306; http://dx.doi.org/10.1113/jphysiol.2007.149633
  • Duarte MS, Paulino PV, Das AK, Wei S, Serão NV, Fu X, Harris SM, Dodson MV, Du M. Enhancement of adipogenesis and fibrogenesis in skeletal muscle of Wagyu compared with Angus cattle. J Anim Sci 2013; 91:2938-46; PMID:23508025; http://dx.doi.org/10.2527/jas.2012-5892
  • Poulos SP, Dodson MV, Hausman GJ. Cell line models for differentiation: preadipocytes and adipocytes. Exp Biol Med (Maywood) 2010; 235:1185-93; PMID:20864461; http://dx.doi.org/10.1258/ebm.2010.010063
  • Hausman GJ, Poulos S. Recruitment and differentiation of intramuscular preadipocytes in stromal-vascular cell cultures derived from neonatal pig semitendinosus muscles. J Anim Sci 2004; 82:429-37; PMID:14974540
  • Hausman GJ, Poulos SP. A method to establish co-cultures of myotubes and preadipocytes from collagenase digested neonatal pig semitendinosus muscles. J Anim Sci 2005; 83:1010-6; PMID:15827245
  • Guo Y, Mo D, Zhang Y, Zhang Y, Cong P, Xiao S, He Z, Liu X, Chen Y. MicroRNAome comparison between intramuscular and subcutaneous vascular stem cell adipogenesis. PLoS One 2012; 7:e45410; PMID:23028990; http://dx.doi.org/10.1371/journal.pone.0045410
  • Grant AC, Ortiz-Colón G, Doumit ME, Tempelman RJ, Buskirk DD. Differentiation of bovine intramuscular and subcutaneous stromal-vascular cells exposed to dexamethasone and troglitazone. J Anim Sci 2008; 86:2531-8; PMID:18539836; http://dx.doi.org/10.2527/jas.2008-0860
  • Poulos SP, Hausman GJ. A comparison of thiazolidinedione-induced adipogenesis and myogenesis in stromal-vascular cells from subcutaneous adipose tissue or semitendinosus muscle of postnatal pigs. J Anim Sci 2006; 84:1076-82; PMID:16612009
  • Du M, Tong J, Zhao J, Underwood KR, Zhu M, Ford SP, Nathanielsz PW. Fetal programming of skeletal muscle development in ruminant animals. J Anim Sci 2010; 88(Suppl):E51-60; PMID:19717774; http://dx.doi.org/10.2527/jas.2009-2311
  • Hausman GJ, Poulos SP. Adipose tissue development in extramuscular and intramuscular depots. Applied Muscle Biology and Meat Science. Boca Raton, FL: CRC Press; 2009. p. 67-81.
  • Allen CE. Cellularity of adipose tissue in meat animals. Fed Proc 1976; 35:2302-7; PMID:782925
  • Gardan D, Gondret F, Louveau I. Lipid metabolism and secretory function of porcine intramuscular adipocytes compared with subcutaneous and perirenal adipocytes. Am J Physiol Endocrinol Metab 2006; 291:E372-80; PMID:16705057; http://dx.doi.org/10.1152/ajpendo.00482.2005
  • Schwab CR, Baas TJ, Stalder KJ, Mabry JW. Deposition rates and accretion patterns of intramuscular fat, loin muscle area, and backfat of Duroc pigs sired by boars from two time periods. J Anim Sci 2007; 85:1540-6; PMID:17296776; http://dx.doi.org/10.2527/jas.2006-343
  • Pannier L, Pethick DW, Geesink GH, Ball AJ, Jacob RH, Gardner GE. Intramuscular fat in the longissimus muscle is reduced in lambs from sires selected for leanness. Meat Sci 2014; 96:1068-75; PMID:23816480; http://dx.doi.org/10.1016/j.meatsci.2013.06.014
  • Castro Bulle FC, Paulino PV, Sanches AC, Sainz RD. Growth, carcass quality, and protein and energy metabolism in beef cattle with different growth potentials and residual feed intakes. J Anim Sci 2007; 85:928-36; PMID:17178805; http://dx.doi.org/10.2527/jas.2006-373
  • Hocquette JF, Gondret F, Baéza E, Médale F, Jurie C, Pethick DW. Intramuscular fat content in meat-producing animals: development, genetic and nutritional control, and identification of putative markers. Animal 2010; 4:303-19; PMID:22443885; http://dx.doi.org/10.1017/S1751731109991091
  • Pethick DW, Harper GS, Oddy VH. Growth, development and nutritional manipulation of marbling in cattle: a review. Aust J Exp Agric 2004; 44:705-15; http://dx.doi.org/10.1071/EA02165
  • Harper GS, Pethick DW. How might marbling begin? Aust J Exp Agric 2004; 44:653-62; http://dx.doi.org/10.1071/EA02114
  • Pickworth CL, Loerch SC, Fluharty FL. Restriction of vitamin A and D in beef cattle finishing diets on feedlot performance and adipose accretion. J Anim Sci 2012; 90:1866-78; PMID:22178850; http://dx.doi.org/10.2527/jas.2010-3590
  • Gorocica-Buenfil MA, Fluharty FL, Reynolds CK, Loerch SC. Effect of dietary vitamin A concentration and roasted soybean inclusion on marbling, adipose cellularity, and fatty acid composition of beef. J Anim Sci 2007; 85:2230-42; PMID:17468427; http://dx.doi.org/10.2527/jas.2006-780
  • Essén-Gustavsson B, Karlsson A, Lundström K, Enfält AC. Intramuscular fat and muscle fibre lipid contents in halothane-gene-free pigs fed high or low protein diets and its relation to meat quality. Meat Sci 1994; 38:269-77; PMID:22059664; http://dx.doi.org/10.1016/0309-1740(94)90116-3
  • Schoonmaker JP, Fluharty FL, Loerch SC. Effect of source and amount of energy and rate of growth in the growing phase on adipocyte cellularity and lipogenic enzyme activity in the intramuscular and subcutaneous fat depots of Holstein steers. J Anim Sci 2004; 82:137-48; PMID:14753357
  • Heyer A, Lebret B. Compensatory growth response in pigs: effects on growth performance, composition of weight gain at carcass and muscle levels, and meat quality. J Anim Sci 2007; 85:769-78; PMID:17296780; http://dx.doi.org/10.2527/jas.2006-164
  • Bee G, Calderini M, Biolley C, Guex G, Herzog W, Lindemann MD. Changes in the histochemical properties and meat quality traits of porcine muscles during the growing-finishing period as affected by feed restriction, slaughter age, or slaughter weight. J Anim Sci 2007; 85:1030-45; PMID:17178814; http://dx.doi.org/10.2527/jas.2006-496
  • Gondret F, Lebret B. Feeding intensity and dietary protein level affect adipocyte cellularity and lipogenic capacity of muscle homogenates in growing pigs, without modification of the expression of sterol regulatory element binding protein. J Anim Sci 2002; 80:3184-93; PMID:12542159
  • Guillerm-Regost C, Louveau I, Sébert SP, Damon M, Champ MM, Gondret F. Cellular and biochemical features of skeletal muscle in obese Yucatan minipigs. Obesity (Silver Spring) 2006; 14:1700-7; PMID:17062798; http://dx.doi.org/10.1038/oby.2006.195
  • He ML, Sharma R, Mir PS, Okine E, Dodson MV. Feed withdrawal abate regimens lipodystrophy and metabolic syndrome symptoms, such as glucose tolerance, are associated with the diameter of retroperitoneal adipocytes in rats. Nutr Res 2010; 30:125-33; PMID:20226998; http://dx.doi.org/10.1016/j.nutres.2009.09.009
  • Mir PS, He ML, Schwartzkopf-Genswein K, Sharma R, Brown FA, Travis G, et al. Effect of supplementation of beef steer diets with oil containing n6 and n3 fatty acids and 48 h feed withdrawal treatments on plasma hormone profiles and adipose tissue cellularity. Livest Sci 2012; 146:140-8; http://dx.doi.org/10.1016/j.livsci.2012.03.001
  • Sørensen MT, Oksbjerg N, Agergaard N, Petersen JS. Tissue deposition rates in relation to muscle fibre and fat cell characteristics in lean female pigs (Sus scrofa) following treatment with porcine growth hormone (pGH). Comp Biochem Physiol A Physiol 1996; 113:91-6; PMID:8624908; http://dx.doi.org/10.1016/0300-9629(95)02038-1
  • Gondret F, Lefaucheur L, Juin H, Louveau I, Lebret B. Low birth weight is associated with enlarged muscle fiber area and impaired meat tenderness of the longissimus muscle in pigs. J Anim Sci 2006; 84:93-103; PMID:16361495
  • Harbison SA, Goll DE, Parrish FC Jr., Wang V, Kline EA. Muscle growth in two genetically different lines of swine. Growth 1976; 40:253-83; PMID:976769
  • Seideman SC, Crouse JD, Mersmann HJ. Carcass, muscle and meat characteristics of lean and obese pigs. J Anim Sci 1989; 67:2950-5; PMID:2592282
  • Mourot J, Kouba M. Development of intra- and intermuscular adipose tissue in growing large white and Meishan pigs. Reprod Nutr Dev 1999; 39:125-32; PMID:10222503; http://dx.doi.org/10.1051/rnd:19990145
  • Clark SL, Wander RC, Hu CY. The effect of porcine somatotropin supplementation in pigs on the lipid profile of subcutaneous and intermuscular adipose tissue and longissimus muscle. J Anim Sci 1992; 70:3435-42; PMID:1459904.
  • Poulos S, Hausman G. Intramuscular adipocytes-potential to prevent lipotoxicity in skeletal muscle. Adipocytes. 2005; 1:79-94
  • Doran O, Moule SK, Teye GA, Whittington FM, Hallett KG, Wood JD. A reduced protein diet induces stearoyl-CoA desaturase protein expression in pig muscle but not in subcutaneous adipose tissue: relationship with intramuscular lipid formation. Br J Nutr 2006; 95:609-17; PMID:16512947; http://dx.doi.org/10.1079/BJN20051526
  • May SG, Savell JW, Lunt DK, Wilson JJ, Laurenz JC, Smith SB. Evidence for preadipocyte proliferation during culture of subcutaneous and intramuscular adipose tissues from Angus and Wagyu crossbred steers. J Anim Sci 1994; 72:3110-7; PMID:7759359
  • Clark BA, Alloosh M, Wenzel JW, Sturek M, Kostrominova TY. Effect of diet-induced obesity and metabolic syndrome on skeletal muscles of Ossabaw miniature swine. Am J Physiol Endocrinol Metab 2011; 300:E848-57; PMID:21304063; http://dx.doi.org/10.1152/ajpendo.00534.2010
  • Reiter SS, Halsey CHC, Stronach BM, Bartosh JL, Owsley WF, Bergen WG. Lipid metabolism related gene-expression profiling in liver, skeletal muscle and adipose tissue in crossbred Duroc and Pietrain Pigs. Comp Biochem Physiol Part D Genomics Proteomics 2007; 2:200-6; PMID:20483293; http://dx.doi.org/10.1016/j.cbd.2007.04.008
  • Gandolfi G, Mazzoni M, Zambonelli P, Lalatta-Costerbosa G, Tronca A, Russo V, Davoli R. Perilipin 1 and perilipin 2 protein localization and gene expression study in skeletal muscles of European cross-breed pigs with different intramuscular fat contents. Meat Sci 2011; 88:631-7; PMID:21420243; http://dx.doi.org/10.1016/j.meatsci.2011.02.020
  • Hausman GJ, Dodson MV, Ajuwon K, Azain M, Barnes KM, Guan LL, Jiang Z, Poulos SP, Sainz RD, Smith S, et al. Board-invited review: the biology and regulation of preadipocytes and adipocytes in meat animals. J Anim Sci 2009; 87:1218-46; PMID:18849378; http://dx.doi.org/10.2527/jas.2008-1427
  • Dodson MV, Jiang Z, Chen J, Hausman GJ, Guan LL, Novakofski J, Thompson DP, Lorenzen CL, Fernyhough ME, Mir PS, et al. Allied industry approaches to alter intramuscular fat content and composition in beef animals. J Food Sci 2010; 75:R1-8; PMID:20492190; http://dx.doi.org/10.1111/j.1750-3841.2009.01396.x
  • Komolka K, Albrecht E, Wimmers K, Michal JJ, Maak S. Molecular Heterogeneities of Adipose Depots-Poten-tial Effects on Adipose-Muscle Cross-Talk in Humans, Mice and Farm Animals. J Genomics 2014; 2:31-44; http://dx.doi.org/10.7150/jgen.5260
  • Dodson MV, Mir PS, Hausman GJ, Guan LL, Du M, Jiang Z, Fernyhough ME, Bergen WG. Obesity, metabolic syndrome, and adipocytes. J Lipids 2011; 2011:721686; PMID:21811683; http://dx.doi.org/10.1155/2011/721686
  • Gao SZ, Zhao SM. Physiology, affecting factors and strategies for control of pig meat intramuscular fat. Recent Pat Food Nutr Agric 2009; 1:59-74; PMID:20653527
  • Jiang Z, Michal JJ, Tobey DJ, Daniels TF, Rule DC, Macneil MD. Significant associations of stearoyl-CoA desaturase (SCD1) gene with fat deposition and composition in skeletal muscle. Int J Biol Sci 2008; 4:345-51; PMID:18825276; http://dx.doi.org/10.7150/ijbs.4.345
  • Wang W, Xue W, Jin B, Zhang X, Ma F, Xu X. Candidate gene expression affects intramuscular fat content and fatty acid composition in pigs. J Appl Genet 2013; 54:113-8; PMID:23275256; http://dx.doi.org/10.1007/s13353-012-0131-z
  • Muoio DM, Koves TR. Skeletal muscle adaptation to fatty acid depends on coordinated actions of the PPARs and PGC1 alpha: implications for metabolic disease. Appl Physiol Nutr Metab 2007; 32:874-83; PMID:18059612; http://dx.doi.org/10.1139/H07-083
  • Cui HX, Liu RR, Zhao GP, Zheng MQ, Chen JL, Wen J. Identification of differentially expressed genes and pathways for intramuscular fat deposition in pectoralis major tissues of fast-and slow-growing chickens. BMC Genomics 2012; 13:213; PMID:22646994; http://dx.doi.org/10.1186/1471-2164-13-213
  • Wu T, Zhang Z, Yuan Z, Lo LJ, Chen J, Wang Y, Peng J. Distinctive genes determine different intramuscular fat and muscle fiber ratios of the longissimus dorsi muscles in Jinhua and landrace pigs. PLoS One 2013; 8:e53181; PMID:23301040; http://dx.doi.org/10.1371/journal.pone.0053181
  • Perera RJ, Marcusson EG, Koo S, Kang X, Kim Y, White N, Dean NM. Identification of novel PPARgamma target genes in primary human adipocytes. Gene 2006; 369:90-9; PMID:16380219; http://dx.doi.org/10.1016/j.gene.2005.10.021
  • Wang Q, Ji C, Huang J, Yang F, Zhang H, Liu L, Yin J. The mRNA of lipin1 and its isoforms are differently expressed in the longissimus dorsi muscle of obese and lean pigs. Mol Biol Rep 2011; 38:319-25; PMID:20358298; http://dx.doi.org/10.1007/s11033-010-0110-6
  • Bosma M, Hesselink MK, Sparks LM, Timmers S, Ferraz MJ, Mattijssen F, van Beurden D, Schaart G, de Baets MH, Verheyen FK, et al. Perilipin 2 improves insulin sensitivity in skeletal muscle despite elevated intramuscular lipid levels. Diabetes 2012; 61:2679-90; PMID:22807032; http://dx.doi.org/10.2337/db11-1402
  • Keady SM, Kenny DA, Ohlendieck K, Doyle S, Keane MG, Waters SM. Proteomic profiling of bovine M. longissimus lumborum from Crossbred Aberdeen Angus and Belgian Blue sired steers varying in genetic merit for carcass weight. J Anim Sci 2013; 91:654-65; PMID:23307841; http://dx.doi.org/10.2527/jas.2012-5850
  • Huuskonen A, Lappalainen J, Oksala N, Santtila M, Häkkinen K, Kyröläinen H, Atalay M. Common genetic variation in the IGF1 associates with maximal force output. Med Sci Sports Exerc 2011; 43:2368-74; PMID:21552154; http://dx.doi.org/10.1249/MSS.0b013e3182220179
  • Ramayo-Caldas Y, Mach N, Esteve-Codina A, Corominas J, Castelló A, Ballester M, Estellé J, Ibáñez-Escriche N, Fernández AI, Pérez-Enciso M, et al. Liver transcriptome profile in pigs with extreme phenotypes of intramuscular fatty acid composition. BMC Genomics 2012; 13:547; PMID:23051667; http://dx.doi.org/10.1186/1471-2164-13-547
  • Chung S, Lapoint K, Martinez K, Kennedy A, Boysen Sandberg M, McIntosh MK. Preadipocytes mediate lipopolysaccharide-induced inflammation and insulin resistance in primary cultures of newly differentiated human adipocytes. Endocrinology 2006; 147:5340-51; PMID:16873530; http://dx.doi.org/10.1210/en.2006-0536
  • Frayling TM, Timpson NJ, Weedon MN, Zeggini E, Freathy RM, Lindgren CM, Perry JR, Elliott KS, Lango H, Rayner NW, et al. A common variant in the FTO gene is associated with body mass index and predisposes to childhood and adult obesity. Science 2007; 316:889-94; PMID:17434869; http://dx.doi.org/10.1126/science.1141634
  • Fan B, Du ZQ, Rothschild MF. The fat mass and obesity-associated (FTO) gene is associated with intramuscular fat content and growth rate in the pig. Anim Biotechnol 2009; 20:58-70; PMID:19370455; http://dx.doi.org/10.1080/10495390902800792
  • Michaud EJ, Mynatt RL, Miltenberger RJ, Klebig ML, Wilkinson JE, Zemel MB, Wilkison WO, Woychik RP. Role of the agouti gene in obesity. J Endocrinol 1997; 155:207-9; PMID:9415049; http://dx.doi.org/10.1677/joe.0.1550207
  • Yabuuchi M, Bando K, Hiramatsu M, Takahashi S, Takeuchi S. Local agouti signaling protein/melanocortin signaling system that possibly regulates lipid metabolism in adipose tissues of chickens. J Poult Sci. 2010;1002010046.
  • Albrecht E, Komolka K, Kuzinski J, Maak S. Agouti revisited: transcript quantification of the ASIP gene in bovine tissues related to protein expression and localization. PLoS One 2012; 7:e35282; PMID:22530003; http://dx.doi.org/10.1371/journal.pone.0035282
  • Bong JJ, Cho KK, Baik M. Comparison of gene expression profiling between bovine subcutaneous and intramuscular adipose tissues by serial analysis of gene expression. Cell Biol Int 2010; 34:125-33; PMID:19947932
  • Kim NK, Park HR, Lee HC, Yoon D, Son ES, Kim YS, Kim SR, Kim OH, Lee CS. Comparative studies of skeletal muscle proteome and transcriptome profilings between pig breeds. Mamm Genome 2010; 21:307-19; PMID:20532784; http://dx.doi.org/10.1007/s00335-010-9264-8
  • Lee SH, Gondro C, van der Werf J, Kim NK, Lim DJ, Park EW, Oh SJ, Gibson JP, Thompson JM. Use of a bovine genome array to identify new biological pathways for beef marbling in Hanwoo (Korean Cattle). BMC Genomics 2010; 11:623; PMID:21062493; http://dx.doi.org/10.1186/1471-2164-11-623
  • Wang YH, Bower NI, Reverter A, Tan SH, De Jager N, Wang R, McWilliam SM, Cafe LM, Greenwood PL, Lehnert SA. Gene expression patterns during intramuscular fat development in cattle. J Anim Sci 2009; 87:119-30; PMID:18820161; http://dx.doi.org/10.2527/jas.2008-1082
  • Albrecht E, Gotoh T, Ebara F, Wegner J, Maak S. Technical note: Determination of cell-specific gene expression in bovine skeletal muscle tissue using laser microdissection and reverse-transcription quantitative polymerase chain reaction. J Anim Sci 2011; 89:4339-43; PMID:21821804; http://dx.doi.org/10.2527/jas.2011-4039
  • Liu R, Sun Y, Zhao G, Wang F, Wu D, Zheng M, Chen J, Zhang L, Hu Y, Wen J. Genome-wide association study identifies Loci and candidate genes for body composition and meat quality traits in Beijing-You chickens. PLoS One 2013; 8:e61172; PMID:23637794; http://dx.doi.org/10.1371/journal.pone.0061172
  • Ramayo-Caldas Y, Mercadé A, Castelló A, Yang B, Rodríguez C, Alves E, Díaz I, Ibáñez-Escriche N, Noguera JL, Pérez-Enciso M, et al. Genome-wide association study for intramuscular fatty acid composition in an Iberian × Landrace cross. J Anim Sci 2012; 90:2883-93; PMID:22785162; http://dx.doi.org/10.2527/jas.2011-4900
  • Murgiano L, D’Alessandro A, Egidi MG, Crisà A, Prosperini G, Timperio AM, Valentini A, Zolla L. Proteomics and transcriptomics investigation on longissimus muscles in Large White and Casertana pig breeds. J Proteome Res 2010; 9:6450-66; PMID:20968299; http://dx.doi.org/10.1021/pr100693h
  • Rajesh RV, Heo GN, Park MR, Nam JS, Kim NK, Yoon D, Kim TH, Lee HJ. Proteomic analysis of bovine omental, subcutaneous and intramuscular preadipocytes during in vitro adipogenic differentiation. Comp Biochem Physiol Part D Genomics Proteomics 2010; 5:234-44; PMID:20656571; http://dx.doi.org/10.1016/j.cbd.2010.06.004
  • Gondret F, Guitton N, Guillerm-Regost C, Louveau I. Regional differences in porcine adipocytes isolated from skeletal muscle and adipose tissues as identified by a proteomic approach. J Anim Sci 2008; 86:2115-25; PMID:18310487; http://dx.doi.org/10.2527/jas.2007-0750
  • Cafe LM, McIntyre BL, Robinson DL, Geesink GH, Barendse W, Greenwood PL. Production and processing studies on calpain-system gene markers for tenderness in Brahman cattle: 1. Growth, efficiency, temperament, and carcass characteristics. J Anim Sci 2010; 88:3047-58; PMID:20525933; http://dx.doi.org/10.2527/jas.2009-2678
  • Wu XX, Yang ZP, Shi XK, Li JY, Ji DJ, Mao YJ, Chang LL, Gao HJ. Association of SCD1 and DGAT1 SNPs with the intramuscular fat traits in Chinese Simmental cattle and their distribution in eight Chinese cattle breeds. Mol Biol Rep 2012; 39:1065-71; PMID:21607624; http://dx.doi.org/10.1007/s11033-011-0832-0
  • Wu Y, Kim JY, Zhou S, Smas CM. Differential screening identifies transcripts with depot-dependent expression in white adipose tissues. BMC Genomics 2008; 9:397; PMID:18721461; http://dx.doi.org/10.1186/1471-2164-9-397
  • Uleberg E, Widerøe IS, Grindflek E, Szyda J, Lien S, Meuwissen TH. Fine mapping of a QTL for intramuscular fat on porcine chromosome 6 using combined linkage and linkage disequilibrium mapping. J Anim Breed Genet 2005; 122:1-6; PMID:16130482; http://dx.doi.org/10.1111/j.1439-0388.2004.00496.x
  • Edwards DB, Ernst CW, Raney NE, Doumit ME, Hoge MD, Bates RO. Quantitative trait locus mapping in an F2 Duroc x Pietrain resource population: II. Carcass and meat quality traits. J Anim Sci 2008; 86:254-66; PMID:17965326; http://dx.doi.org/10.2527/jas.2006-626
  • Quintanilla R, Pena RN, Gallardo D, Cánovas A, Ramírez O, Díaz I, Noguera JL, Amills M. Porcine intramuscular fat content and composition are regulated by quantitative trait loci with muscle-specific effects. J Anim Sci 2011; 89:2963-71; PMID:21571897; http://dx.doi.org/10.2527/jas.2011-3974
  • Wang H, Xiong K, Sun W, Fu Y, Jiang Z, Yu D, Liu H, Chen J. Two completely linked polymorphisms in the PPARG transcriptional regulatory region significantly affect gene expression and intramuscular fat deposition in the longissimus dorsi muscle of Erhualian pigs. Anim Genet 2013; 44:458-62; PMID:23402337; http://dx.doi.org/10.1111/age.12025
  • Barendse W. Haplotype analysis improved evidence for candidate genes for intramuscular fat percentage from a genome wide association study of cattle. PLoS One 2011; 6:e29601; PMID:22216329; http://dx.doi.org/10.1371/journal.pone.0029601
  • Bolormaa S, Neto LR, Zhang YD, Bunch RJ, Harrison BE, Goddard ME, Barendse W. A genome-wide association study of meat and carcass traits in Australian cattle. J Anim Sci 2011; 89:2297-309; PMID:21421834; http://dx.doi.org/10.2527/jas.2010-3138
  • Xie H, Lim B, Lodish HF. MicroRNAs induced during adipogenesis that accelerate fat cell development are downregulated in obesity. Diabetes 2009; 58:1050-7; PMID:19188425; http://dx.doi.org/10.2337/db08-1299
  • Lin Q, Gao Z, Alarcon RM, Ye J, Yun Z. A role of miR-27 in the regulation of adipogenesis. FEBS J 2009; 276:2348-58; PMID:19348006; http://dx.doi.org/10.1111/j.1742-4658.2009.06967.x
  • Jin W, Grant JR, Stothard P, Moore SS, Guan LL. Characterization of bovine miRNAs by sequencing and bioinformatics analysis. BMC Mol Biol 2009; 10:90; PMID:19758457; http://dx.doi.org/10.1186/1471-2199-10-90
  • Jin W, Dodson MV, Moore SS, Basarab JA, Guan LL. Characterization of microRNA expression in bovine adipose tissues: a potential regulatory mechanism of subcutaneous adipose tissue development. BMC Mol Biol 2010; 11:29; PMID:20423511; http://dx.doi.org/10.1186/1471-2199-11-29
  • Wang H, Zheng Y, Wang G, Li H. Identification of microRNA and bioinformatics target gene analysis in beef cattle intramuscular fat and subcutaneous fat. Mol Biosyst 2013; 9:2154-62; PMID:23728155; http://dx.doi.org/10.1039/c3mb70084d
  • Romao JM, Jin W, He M, McAllister T, Guan LL. Altered microRNA expression in bovine subcutaneous and visceral adipose tissues from cattle under different diet. PLoS One 2012; 7:e40605; PMID:22815773; http://dx.doi.org/10.1371/journal.pone.0040605
  • Sato F, Tsuchiya S, Meltzer SJ, Shimizu K. MicroRNAs and epigenetics. FEBS J 2011; 278:1598-609; PMID:21395977; http://dx.doi.org/10.1111/j.1742-4658.2011.08089.x
  • Albu JB, Kovera AJ, Allen L, Wainwright M, Berk E, Raja-Khan N, Janumala I, Burkey B, Heshka S, Gallagher D. Independent association of insulin resistance with larger amounts of intermuscular adipose tissue and a greater acute insulin response to glucose in African American than in white nondiabetic women. Am J Clin Nutr 2005; 82:1210-7; PMID:16332653
  • Tsintzas K, Chokkalingam K, Jewell K, Norton L, Macdonald IA, Constantin-Teodosiu D. Elevated free fatty acids attenuate the insulin-induced suppression of PDK4 gene expression in human skeletal muscle: potential role of intramuscular long-chain acyl-coenzyme A. J Clin Endocrinol Metab 2007; 92:3967-72; PMID:17652214; http://dx.doi.org/10.1210/jc.2007-1104
  • Dina C, Meyre D, Gallina S, Durand E, Körner A, Jacobson P, Carlsson LM, Kiess W, Vatin V, Lecoeur C, et al. Variation in FTO contributes to childhood obesity and severe adult obesity. Nat Genet 2007; 39:724-6; PMID:17496892; http://dx.doi.org/10.1038/ng2048
  • Scuteri A, Sanna S, Chen WM, Uda M, Albai G, Strait J, Najjar S, Nagaraja R, Orrú M, Usala G, et al. Genome-wide association scan shows genetic variants in the FTO gene are associated with obesity-related traits. PLoS Genet 2007; 3:e115; PMID:17658951; http://dx.doi.org/10.1371/journal.pgen.0030115
  • Kang ES, Park SE, Han SJ, Kim SH, Nam CM, Ahn CW, Cha BS, Kim KS, Lee HC. LPIN1 genetic variation is associated with rosiglitazone response in type 2 diabetic patients. Mol Genet Metab 2008; 95:96-100; PMID:18693052; http://dx.doi.org/10.1016/j.ymgme.2008.06.011

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.