Hostname: page-component-7479d7b7d-pfhbr Total loading time: 0 Render date: 2024-07-11T19:59:32.438Z Has data issue: false hasContentIssue false

Key signalling factors and pathways in the molecular determination of skeletal muscle phenotype

Published online by Cambridge University Press:  17 May 2007

K.C. Chang*
Affiliation:
Molecular Medicine Laboratory, Division of Animal Production and Public Health, University of Glasgow Veterinary School, Bearsden Road, Glasgow, G61 1QH, UK

Abstract

The molecular basis and control of the biochemical and biophysical properties of skeletal muscle, regarded as muscle phenotype, are examined in terms of fibre number, fibre size and fibre types. A host of external factors or stimuli, such as ligand binding and contractile activity, are transduced in muscle into signalling pathways that lead to protein modifications and changes in gene expression which ultimately result in the establishment of the specified phenotype. In skeletal muscle, the key signalling cascades include the Ras-extracellular signal regulated kinase-mitogen activated protein kinase (Erk-MAPK), the phosphatidylinositol 3′-kinase (PI3K)-Akt1, p38 MAPK, and calcineurin pathways. The molecular effects of external factors on these pathways revealed complex interactions and functional overlap. A major challenge in the manipulation of muscle of farm animals lies in the identification of regulatory and target genes that could effect defined and desirable changes in muscle quality and quantity. To this end, recent advances in functional genomics that involve the use of micro-array technology and proteomics are increasingly breaking new ground in furthering our understanding of the molecular determinants of muscle phenotype.

Type
Full Papers
Copyright
Copyright © The Animal Consortium 2007

Introduction

Importance of muscle fibre phenotype

The key constituents of muscle are muscle fibres, and their associated extracellular matrix in which are located the vascular supply, collagenous component and adipose tissue (intramuscular fat) of muscle. Muscle phenotype conferred by fibres is central to the quantity and quality of meat production. Quantity is the outcome of muscle fibre hyperplasia and hypertrophy. Hyperplastic growth in utero is primarily dependent on myocyte proliferation and differentiation (Oka et al., Reference Oka, Iwaki, Dohgo, Ohtagaki, Noda, Shiozaki, Endoh and Ozaki2002). Post-natal growth is mainly the result of hypertrophy or enlargement of existing and replaced muscle fibres. Quality is a more complex trait and is assessed by a variety of objective and subjective measurements, such as colour, pH, tenderness, odour and juiciness. It is well established that fibre type composition is a key determinant of both meat quantity and quality. In the pig, in particular, favourable meat traits such as tenderness and colour, have been found to closely associate with the greater abundance of oxidative fibres (Klont et al., Reference Klont, Brocks and Eikelenboom1998; Karlsson et al., Reference Karlsson, Klont and Fernandez1999; Chang et al., Reference Chang, Da Costa, Blackley, Southwood, Evans, Plastow, Wood and Richardson2003; Maltin et al., Reference Maltin, Balcerzak, Tilley and Delday2003; Wood et al., Reference Wood, Nute, Richardson, Whittington, Southwood, Plastow, Mansbridge, da Costa and Chang2004). Red or highly oxidative muscles possess higher lipid concentration which is associated with more tender meat (Hocquette et al., Reference Hocquette, Ortigues-Marty, Pethick, Herpin and Fernandez1998; Wood et al., Reference Wood, Enser, Fisher, Nute, Richardson and Sheard1999 and Reference Wood, Richardson, Nute, Fisher, Campo, Kasapidou, Sheard and Enser2003). Fast-glycolytic fibres (see later for details), on the other hand, being relatively the largest fibre type are a major fibre type for muscle hypertrophy as well as in the predisposition of pale, soft and exudative (PSE) pork independent of ryanodine receptor 1 mutation.

In the living animal, muscle phenotype is a major factor in the functional integrity of mobility and movement, as well as of thermoregulation, which in turn governs its quality of life. A wide range of disease conditions often adversely affect muscle phenotype and lead to debilitating muscle wasting or atrophy. For instance, in human obesity, triglyceride accumulation in muscle (intramuscular fat) appears to be related to the development of insulin resistance and type 2 diabetes (Morio et al., Reference Morio, Hocquette, Montaurier, Boirie, Bouteloup-Demange, McCormack, Fellmann, Beaufrere and Ritz2001; Kelley et al., Reference Kelley, Goodpaster and Storlien2002; Wolfe, Reference Wolfe2006). Hence, knowledge of the molecular events that affect muscle phenotype is of fundamental, agricultural, welfare and biomedical importance. By understanding the signalling processes and gene targets that effect muscle phenotype, strategic approaches, such as the use of marker-assisted selection (Beuzen et al., Reference Beuzen, Stear and Chang2000), dietary manipulation (Casser-Malek et al., Reference Casser-Malek, Hocquette, Jurie, Listrat, Jailler, Bauchart, Briand and Picard2004) or even pharmacological targeting of novel effector genes, could be developed to improve muscle quality and quantity.

This review endeavours to bring together recent advances in our understanding of the signalling pathways, and molecular mechanisms of several major growth promoting factors that affect muscle fibre phenotype relevant to farm animal production. Given that the subject area covered is wide, its purposes are to provide the reader with a general overview of the molecular aspects of muscle phenotype determination in the live animal and to highlight avenues for future research exploitation.

What is muscle fibre phenotype?

Post-natal muscle is a highly heterogeneous syncytial tissue, comprising muscle fibres and extracellular matrix, with the ability to rapidly undergo biochemical and physical fibre changes in response to appropriate external stimuli, such as nervous and hormonal stimulations, to adapt to the accompanying functional demands imposed on it (Caiozzo, Reference Caiozzo2004). The feature of functional plasticity indicates that skeletal muscle is highly amenable to changes in coordinated gene expression. Although the extracellular matrix of muscle, as exemplified by the orderly arrangement of endomycium, perimycium and epimycium, is a key determinant of meat quality (Fang et al., Reference Fang, Nishimura and Takahashi1999; McCormick, Reference McCormick1999), this review is focussed on the regulation of muscle fibre phenotype. The physical and biochemical characteristics of muscle fibres can be conveniently defined under three distinct but overlapping categories: fibre number, fibre size and fibre type (Figure 1). Fibre types reflect differences in their biochemical (metabolic) and biophysical properties that arise from differences in coordinated expression of muscle gene isoforms (Schiaffino and Reggiani, Reference Schiaffino and Reggiani1996).

Figure 1 Categories of skeletal muscle fibre phenotype.

A feature of muscle fibre phenotype regulation that will be apparent is that the same signalling factor or pathway (such as IGF-1) can often have multiple phenotypic effects with regard to its influence on fibre number, size and type.

Muscle formation and regeneration

In vertebrates, most skeletal muscles are derived from muscle progenitor or precursor cells present in somites, which arise by segmentation of the paraxial mesoderm, located on either side of the notochord and neural tube in the early embryo. As it develops, each somite can be divided into the epaxial dermomyotome, which gives rise to the epaxial muscles, and into the hypaxial dermomyotome, from which derives the rest of the body and limb muscles. The first muscle mass to form, under the dermomyotome, is the myotome, which contributes to the trunk muscles. The early developmental events surrounding myogenesis are under intense molecular scrutiny (Buckingham, Reference Buckingham2001). Myogenic cell fate specification is acquired from the activation of myogenic determination genes, namely the basic helix-loop-helix (bHLH) proteins encoded by Myf-5 and MyoD. The other two members of the myogenic bHLH gene family, myogenin and Mrf4, are implicated, along with MyoD, in the subsequent activation of muscle-specific genes during myogenic differentiation. MEF-2 family, characterised by the presence of a MADS-box motif, also plays an important role in muscle differentiation (Buckingham et al., Reference Buckingham, Bajard, Chang, Daubas, Hadchouel, Meilhac, Montarras, Rocancourt and Relaix2003). Prior to the activation of Myf-5 or MyoD, signallings by Wnts, produced by the dorsal neural tube and surface ectoderm, and hedgehog (Hh) proteins, from the notochord and floor plate of the neural tube, are needed to serve as early triggers or inductive signals of myogenesis (Tajbakhsh and Buckingham, Reference Tajbakhsh and Buckingham2000). Hh signalling plays an additional role in committing progenitor cells to the slow-twitch lineage, a process specified by Blimp-1, originally identified as a factor that promotes B-cell maturation (Baxendale et al., Reference Baxendale, Davison, Muxworthy, Wolff, Ingham and Roy2004).

In limb muscle development, muscle progenitor cells opposite the limb buds must first delaminate from the hypaxial dermomyotome and migrate into the limb region. A number of homeobox proteins have been found to be essential for this migrational process (e.g. Pax3, c-met and its ligand hepatocyte growth factor (HGF)/scatter factor, and Lbx1) (Buckingham et al., Reference Buckingham, Bajard, Chang, Daubas, Hadchouel, Meilhac, Montarras, Rocancourt and Relaix2003). Upon arriving at their destination, they begin to express MyoD and Myf-5 to mark the onset of myogenesis.

After appropriate rounds of cell proliferation, myoblasts leave the cell cycle and fuse to form multinucleated myotubes. The homeobox proteins of Mox2 and Msx1, and several growth factors, including insulin-like growth factor (IGF)-1 and -2 (Morali et al., Reference Morali, Jouneau, McLaughlin, Thiery and Larue2000), fibroblast growth factors (FGFs), bone morphogenic proteins (BMPs) and platelet-derived growth factors, variously contribute to the regulation of proliferation and differentiation (Zorzano et al., Reference Zorzano, Kaliman, Guma and Palacin2003). The first muscle fibres that appear are known as primary fibres (around gestation day 35 in the pig and 12.5 days in the mouse), around which subsequent secondary fast fibres form at the time when innervation begins to be established (beginning at day 50 to day 87 in the pig and day 15 in the mouse) (Picard et al., Reference Picard, Lefaucheur, Berri and Duclos2002; Buckingham et al., Reference Buckingham, Bajard, Chang, Daubas, Hadchouel, Meilhac, Montarras, Rocancourt and Relaix2003; Da Costa et al., Reference Da Costa, McGillivray and Chang2003; Caiozzo, Reference Caiozzo2004).

As post-mitotic cells, skeletal muscle fibres are unable to undergo self-replication. Damaged fibres are replaced with newly synthesised fibres through the proliferation, differentiation and subsequent fusion of muscle satellite cells (myoblasts). Satellite cells constitute a reservoir of undifferentiated muscle precursor cells, located between the basal lamina and the sarcolemma, that are activated in response to muscle injury or growth stimuli to proliferate and ultimately fuse to generate new fibres. Muscle satellite cells are characterised by the expression of myostatin, a transforming growth factor (TGF)-β like factor implicated in limiting muscle growth, c-met, Myf-5 and Pax-7 (Buckingham, Reference Buckingham2001; Buckingham et al., Reference Buckingham, Bajard, Chang, Daubas, Hadchouel, Meilhac, Montarras, Rocancourt and Relaix2003). It appears that not all muscle satellite cells are derived from somitic cells. There exists other population(s) of muscle precursor cells in skeletal muscle, the adult muscle stem cells, with haematopoietic and myogenic potentials (Asakura, Reference Asakura2003; Relaix, Reference Relaix2006). These adult muscle stem cells express haematopoietic markers, such as c-kit, CD45, CD34 and Sca1 (markers that are absent on satellite cells) but not myogenic markers like Pax7 and Myf-5 (present in satellite cells) (Asakura, Reference Asakura2003; Polesskaya et al., Reference Polesskaya, Seale and Rudnicki2003). Recent work suggests that during growth and regeneration, adult stem cells proliferate and undergo phenotypic conversion, which involves Wnt signalling, into myogenic satellite cells (Asakura, Reference Asakura2003; Polesskaya et al., Reference Polesskaya, Seale and Rudnicki2003). It is increasingly apparent that many of the regulatory genes involved in embryonic myogenesis are also required in post-natal muscle fibre formation by satellite cells.

Fibre number: mediators of muscle hyperplasia

It is generally regarded that by birth, an animal, such as the pig, would have nearly the same number of muscle fibres as in adulthood (Wigmore and Stickland, Reference Wigmore and Stickland1983). Therefore, the extent of fibre number formation or hyperplasia during foetal development will have a major bearing on muscle growth potential. In post-natal muscle, fibre number may not necessarily be constant. Periodic repair and replacement of damaged fibres are necessary to maintain functional integrity, a process performed by satellite cells that can proliferate and fuse with damaged fibres or fuse to form new fibres (Goldring et al., Reference Goldring, Partridge and Watt2002). The relative contribution of satellite cells to the formation of new fibres (hyperplastic growth) and to existing fibres (hypertrophic growth) is not clear, but is likely to be dependent on the nature of the inductive signals. Fibre number determination can be affected by any factor that plays a role in embryonic or post-natal myogenesis through myoblast specification, proliferation and/or differentiation. The main signalling factors to be considered that affect fibre number are IGF-1 and -2, myostatin and nutrition.

IGF-1 and IGF-2 activate Erk-MAPK pathway of cell proliferation (hyperplasia)

Growth factors such as IGFs and FGF by virtue of their ability to stimulate cell proliferation are regarded as potent agents that can affect fibre number, pre- and post-natally (Bass et al., Reference Bass, Oldham, Sharma and Kambadur1999). Indeed, daily injections of growth hormone (GH) in the sow during early pregnancy has been shown to enhance foetal fibre number (Rehfeldt et al., Reference Rehfeldt, Kuhn, Vanselow, Fürbass, Fiedler, Nürnberg, Clelland, Stickland and Ender2001), an effect likely to be mediated by IGF-1. IGF-1 and -2 stimulate both muscle cell proliferation and differentiation through the interaction with the type 1 IGF receptor (IGF1R), insulin receptor exon 11- (IR-A) and insulin receptor exon11+ (IR-B), all of which are transmembrane tyrosine kinase receptors (Denley et al., Reference Denley, Cosgrove, Booker, Wallace and Forbes2005). These activated receptors initiate two major signalling cascades (Figures 2 and 3). The type 2 IGF receptor (IGF-2R) (known also as mannose-6-phosphate receptor) and a family of high affinity IGF binding proteins (IGFBPs) 1 to 6 modulate the availability of IGF-1 and IGF2 to bind to receptors. IGF-2R has no intrinsic signalling transduction capability and serves to sequester IGF-2 from potential receptor interactions and to internalise and degrade IGF-2 (Denley et al., Reference Denley, Cosgrove, Booker, Wallace and Forbes2005). Liver is the main endocrine source of IGF-1 (IGF-1Ea), which is induced by GH. Additionally, damaged, stretched or load-bearing muscles are local sources IGF-1 and IGF-2 which induce proliferation and hypertrophy in an autocrine / paracrine manner (Adams, Reference Adams2002). Besides the major IGF-1Ea isoform, a spliced variant with a carboxyl terminus different from IGF-1Ea, named MGF (mechano growth factor), is rapidly inducible in skeletal muscle and appears to be an early trigger for satellite cell proliferation (Yang and Goldspink, Reference Yang and Goldspink2002). The autocrine / paracrine muscle production of IGF-2 during muscle differentiation and regeneration was found to participate in a positive feedback loop to further enhance muscle differentiation (Erbay et al., Reference Erbay, Park, Nuzzi, Schoenherr and Chen2003; Wilson et al., Reference Wilson, Hsieh and Rotwein2003) and regeneration (Kirk et al., Reference Kirk, Oldham, Jeanplong and Bass2003). Although both IGF-1 and -2 are clearly important to muscle growth, their relative contribution to cell proliferation, differentiation and hypertrophy is uncertain.

Figure 2 The IGF-induced Erk-MAPK proliferation signalling pathway of hyperplasia. IGF-1, IGF-2 or insulin binding to tyrosine kinase receptor triggers a phosphorylation cascade that leads to transcriptional and protein modifications, culminating in increase cell proliferation. Integrin can also activate the Erk-MAPK pathway independent of tyrosine kinase receptor binding. Note that activated Ras can also activate the PI3K pathway.

Figure 3 The IGF-induced PI3K-Akt1 signalling pathway of muscle differentiation and hypertrophy. It affects a range of cellular processes, including the promotion of protein synthesis, glycogen synthesis, and cell proliferation and differentiation, that culminate in muscle hypertrophy. mTOR can be activated by amino acids (nutrients) and induce local IGF-2 production which, in turn, acts in a positive feedback manner. The PI3K cascade can be independently activated by exercise and by integrin stimulation (see also integrin stimulation of Erk-MAPK pathway, Figure 2). Activated Akt1 via inhibiting Raf and MEK phosphorylation is involved in inhibitory crosstalk with the Erk-MAPK pathway. Selected abbreviations: PI3K =  phosphatidylinositol 3′-kinase, PDK1 =  phosphatidylinositol-dependent protein kinase 1, GSK3β =  glycogen synthase kinase 3 β, mTOR =  mammalian target of rapamycin, GS =  glycogen synthase, and p70S6K =  ribosomal protein S6 kinase. Note that Akt, GSK3β, mTOR, p70S6k and PHAS-1 are serine/threonine kinases.

IGF-1 (as well as insulin and IGF-2) binding activates the receptor tyrosine kinase (IGF1R or insulin receptor), which subsequently recruits insulin receptor substrate 1 (IRS-1), a non-enzymatic docking protein that propagates the signal to two crucial signalling pathways, the mitogen-activated protein kinase (Erk-MAPK) pathway, via Ras-Raf-MEK-Erk (Rommel et al., Reference Rommel, Clarke, Zimmermann, Nuñez, Rossman, Reid, Moelling, Yancopoulos and Glass1999; Zimmermann and Moelling, Reference Zimmermann and Moelling1999; Glass, Reference Glass2003a) and the phosphatidylinositol 3’-kinase (PI3K)-Akt1 pathway (Glass, Reference Glass2003b; Vollenweider, Reference Vollenweider2003). The Ras-Erk MAP kinase (Erk-MAPK) pathway, a serine/threonine phosphorylation cascade, is responsible for cell proliferation and plays an important role in hyperplastic growth (Figure 2). The PI3K-Akt1 pathway is a major route to muscle differentiation and hypertrophy (see later section). Ras is a membrane bound GTPase, and cycles between an inactive Ras-GDP to an active Ras-GTP. Expression of oncogenic Ras inhibits myogenic differentiation (Mitin et al., Reference Mitin, Kudla, Konieczny and Taparowsky2001) in part by promoting cell proliferation via the MAP kinase pathway. Interestingly, activated Ras was also found to induce the expression of slow myosin heavy chain (MyHC), although the downstream mechanism that leads to this fibre type-specific effect is not understood (Murgia et al., Reference Murgia, Serrano, Calabria, Pallafacchina, Lømo and Schiaffino2000). Both Erk-MAPK and PI3K-Akt1 pathways are necessary and complement each other in mediating post-natal muscle growth (Haddad and Adams, Reference Haddad and Adams2004).

Myostatin inhibits cell proliferation

Myostatin (Mst), also known as growth and differentiation factor 8 (GDF-8), is a secreted negative regulator of muscle mass that belongs to the transforming growth factor (TGF)-β superfamily (Kocamis and Killefer, Reference Kocamis and Killefer2002). Its biological significance is demonstrated in several breeds of double-muscling cattle (e.g. Belgian Blue and Piedmontese) and Mst-null mice as dramatic increase of muscularity through fibre hypertrophy and hyperplasia (Grobet et al., Reference Grobet, Martin, Poncelet, Pirottin, Brouwers, Riquet, Schoeberlein, Dunner, Ménissier, Massabanda, Fries, Hanset and Georges1997; McPherron and Lee, Reference McPherron and Lee1997; Marchitelli et al., Reference Marchitelli, Savarese, Crisa, Nardone, Marsan and Valentini2003). The Belgian Blue carries a naturally occurring homozygous 11-bp deletion in the coding region of Mst whereas the Piedmontese possesses a homozygous G → A point mutation that changes the cysteine residue in Mst to a tyrosine (Kambadur et al., Reference Kambadur, Sharma, Smith and Bass1997). The porcine Mst gene has been molecularly cloned and characterised, but a mutation that confers the enlarged muscle phenotype has not been described (Ji et al., Reference Ji, Losinski, Cornelius, Frank, Willis, Gerrard, Depreux and Spurlock1998). Mst arrests muscle cells in the G1 and G2 phases of the cell cycle, through the up-regulation of cyclin-dependent kinase (cdk) inhibitor p21 and down-regulation of cdk-2, thereby inhibiting cell proliferation (Thomas et al., Reference Thomas, Langley, Berry, Sharma, Kirk, Bass and Kambadur2000). This inhibition is mediated, at least in part, through the p38 MAPK stress response pathway (Philip et al., Reference Philip, Lu and Gao2005) (Figure 4). Mst also appears to directly inhibit muscle differentiation by interfering with the activity of MyoD (Langley et al., Reference Langley, Thomas, Bishop, Sharma, Gilmour and Kambadur2002). Muscle wasting conditions, such through disease or disuse atrophy, are associated with elevated Mst in affected muscles (McCroskery et al., Reference McCroskery, Thomas, Maxwell, Sharma and Kambadur2003). A primary effect of Mst inactivation in post-natal growth is an increase of satellite cell proliferation and differentiation (McCroskery et al., Reference McCroskery, Thomas, Maxwell, Sharma and Kambadur2003). GH, mediated through IGF-1, was shown to inhibit Mst expression in skeletal muscles and C2C12 myotubes (Liu et al., Reference Liu, Thomas, Asa, Gonzalez-Cadavid, Bhasin and Ezzat2003). Conversely, Mst expression in porcine embryonic myogenic cells up-regulates the production of IGF binding protein-3 (IGFBP-3), which reduces the activities of IGF-1 and -2 (Kamanga-Sollo et al., Reference Kamanga-Sollo, Pampusch, White and Dayton2003). These results suggest that the Mst and IGF signalling pathways are closely connected in an antagonistic manner.

Figure 4 The p38 MAPK stress response pathway of cell arrest and atrophy. It is activated by a variety of stimuli, including exercise and activated Ras, and through different receptor families. It mediates a range of effects: cell arrest, differentiation, atrophy as well as mitochondrial biogenesis. Additionally, TGFβ receptors readily signal through Smad proteins which also lead to cell arrest. Ski functions to inhibit Smad signalling thus relieving cell arrest and promoting proliferation. Selected abbreviations: IL-1 =  interleukin 1, TLRs =  toll-like receptors, TNF α =  tumour necrosis factor α, and TAK =  TGFβ-activated kinase.

Recent micro-array investigations further found that Mst-null mice exhibited raised Wnt4 expression which stimulated satellite cell proliferation (Steelman et al., Reference Steelman, Recknor, Nettleton and Reecy2006). The use of both micro-array and proteomic analyses showed a clear fibre phenotype switch in Mst-null muscles from slow to fast-twitch fibres (Bouley et al., Reference Bouley, Meunier, Chambon, De Smet, Hocquette and Picard2006; Steelman et al., Reference Steelman, Recknor, Nettleton and Reecy2006) (see section IGF-1, β2-agonist and myostatin-null signal fast fibre phenotype). Mst has been recently found to promote adipogenesis in C3H 10T(1/2) cells which appears to be associated with adipocyte lineage commitment (Artaza et al., Reference Artaza, Bhasin, Magee, Reisz-Porszasz, Shen, Groome, Fareez and Gonzalez-Cadavid2005). However, in committed bovine preadipocytes, Mst clearly suppresses differentiation of preadipocytes into adipocytes (Hirai et al., Reference Hirai, Matsumoto, Hino, Kawachi, Matsui and Yano2007).

Nutrition on fibre number and characteristics

Nutrition in animal growth is a complex subject of major importance, which due to its enormity cannot be fully addressed in the present review. Maternal nutrition is undoubtedly a key factor in foetal growth and its subsequent survival which manifests its effects through a complex signalling network that includes IGFs (Hornick et al., Reference Hornick, Van Eenaeme, Gérard, Dufrasne and Istasse2000; Maak et al., Reference Maak, Jaesert, Neumann, Yerle and Von Lengerken2001; Wu et al., Reference Wu, Bazer, Wallace and Spencer2006). However, in the context of foetal fibre number, the precise role of maternal nutrition is not entirely clear. Earlier work suggests that doubling food intake during early sow pregnancy increases the number of secondary fibres in the newborn, which would confer greater potential for post-natal growth (Dwyer et al., Reference Dwyer, Stickland and Fletcher1994). More recent work, however, could not reproduce this finding (Nissen et al., Reference Nissen, Danielsen, Jorgensen and Oksbjerg2003).

Connected to nutrition, mild dietary restriction on beef cattle has little apparent effect on muscle characteristics or meat quality (Casser-Malek et al., Reference Casser-Malek, Hocquette, Jurie, Listrat, Jailler, Bauchart, Briand and Picard2004). However, dietary changes during weaning in calves appear to have an appreciable effect on increasing the number of oxidative fibres (Picard et al., Reference Picard, Gagniere, Geay, Hocquette and Robelin1995). Under moderate protein and energy dietary restriction (20% less protein and 7% less energy) in young growing pigs, we recently found, with the use of a porcine cDNA muscle micro-array, significant increase in the accumulation of intramuscular fat, which could have production implications on meat quality (Da Costa et al., Reference Da Costa, McGillivray, Bai, Wood, Evans and Chang2004). A similar dietary restriction study has been reported on Brahman cattle but no data on meat characteristics are provided (Byrne et al., Reference Byrne, Wang, Lehnert, Harper, McWilliam, Bruce and Reverter2005). Lamb and beef animals raised on grass showed greater flavour intensity in comparison with grain-fed animals because of a higher accumulation of 18:3 polyunsaturated fatty acids from a grass diet (Wood et al., Reference Wood, Enser, Fisher, Nute, Richardson and Sheard1999).

Fibre size: mediators of muscle hypertrophy/atrophy

Post-natal muscle growth is primarily a function of enlargement and elongation of existing fibres as a result of net protein synthesis, known as muscle hypertrophy (Figure 1) (Glass, Reference Glass2003b). Conversely, loss of muscle mass as a consequence of disease or muscle inactivity is primarily due to a reduction in fibre size, described as muscle atrophy (Glass, Reference Glass2003b). Muscle differentiation is a prerequisite to hypertrophy and, as detailed below, the same factors are often involved in both processes (differentiation and hypertrophy). In this section, we consider some well known signalling factors and pathways that mediate muscle hypertrophy or atrophy (IGF-PI3K pathway, contractile activity, p38 MAP kinase pathway, synthetic β2 adrenergic agonists, anabolic steroids and ski proto-oncogene.

IGF activates PI3K-Akt1 pathway of hypertrophy

The IGF-activated PI3K-Akt1 signalling pathway (Figure 3) is widely regarded as the primary route to skeletal muscle differentiation and hypertrophy (Coolican et al., Reference Coolican, Samuel, Ewton and McWade1997; Jiang et al., Reference Jiang, Zheng and Vogt1998; Glass, Reference Glass2003b). Phosphatidylinositol 3′-kinase (PI3K) is a dimeric lipid kinase that catalyses the phosphorylation of membrane-bound phosphatidylinositol (4,5)-bisphosphate (PIP2) to phosphatidylinositol (3,4,5)-trisphosphate (PIP3). PIP3 provides a membrane-binding site for Akt1 (also known as protein kinase B) a serine/threonine kinase, and phosphatidylinositol-dependent protein kinase 1 (PDK1) (Figure 3). Akt1 is phosphorylated by PDK1, and once activated Akt1 phosphorylates a number of substrates that are responsible for a range of growth processes, which include protein synthesis, glycogen synthesis, muscle differentiation and cell proliferation. A key target of activated Akt1 is the phosphorylation of mTOR (mammalian target of rapamycin), a serine/threonine kinase, which in turn activates a number of downstream effectors including ribosomal protein S6 kinase (p70S6K) and eIF-4E, both of which mediate translation initiation and appear to play important roles in muscle hypertrophy. mTOR has also been shown to mediate myogenesis by a process that is independent of its kinase activity, although the exact mechanism involved is unclear (Erbay and Chen, Reference Erbay and Chen2001). During differentiation, mTOR was found to induce the local production of IGF-2, and it seems that the myogenic effect of mTOR is mediated in part through a positive feedback loop with IGF-2 (Erbay et al., Reference Erbay, Park, Nuzzi, Schoenherr and Chen2003). There is evidence that mTOR can also be activated by amino acids, independent of Akt1 (Parkington et al., Reference Parkington, Siebert, LeBrasseur and Fielding2003) and that muscles of nutrient-restricted pregnant cows showed reduced phosphorylation activation of mTOR (Du et al., Reference Du, Zhu, Means, Hess and Ford2005).

Glycogen synthase kinase 3β (GSK3β), a serine/threonine kinase, is another major substrate of activated Akt1 that modulates muscle hypertrophy (Figure 3). Like mTOR, GSK3β can also be phosphorylated independent of Akt1. Activation of protein kinase A or C is able to lead to the phosphorylation of GSK3β (Jope, Reference Jope2003). Unstimulated (dephosphorylated) GSK3β is active and its phosphorylation activity appears to act as a general repressor on its substrates. Many substrates of GSK3β have been identified, including glycogen synthase, c-Jun, cyclin D1, β-catenin, nuclear factors of activated T-cells (NFATs), and Notch2 (Masuda et al., Reference Masuda, Imamura, Amasaki, Arai and Arai1998; Espinosa et al., Reference Espinosa, Ingles-Esteve, Aguilera and Bigas2003). The inhibition of GSK3β by phosphorylation, leads to dephosphorylation activation of glycogen synthase, resulting in glycogen synthesis, and to activation of eIF-2B, which promotes protein synthesis (Rommel et al., Reference Rommel, Bodine, Clarke, Rossman, Nunez, Stitt, Yancopoulos and Glass2001). A primary effect of GSK3β inhibition on muscle is the promotion of hypertrophy (Rommel et al., Reference Rommel, Bodine, Clarke, Rossman, Nunez, Stitt, Yancopoulos and Glass2001; Vyas et al., Reference Vyas, Spangenburg, Abraha, Childs and Booth2002). Clearly, with its wide range of targets, GSK3β can mediate a host of other effects. GSK3β inhibition, by lithium chloride, improves glucose uptake of muscle cells, which makes it a pharmacological target for the treatment of diabetes (MacAulay et al., Reference MacAulay, Hajduch, Blair, Coghlan, Smith and Hundal2003). Wnt signalling leads to the inhibition of GSK3β, an integral component of the linear cascade that results in the stabilisation of β-catenin, a co-activator necessary for cell proliferation (Novak and Dedhar, Reference Novak and Dedhar1999; Giles et al., Reference Giles, Van Es and Clevers2003) and myogenic differentiation (Martin et al., Reference Martin, Schneider, Janetzky, Waibler, Pandur, Kuhl, Behrens, der Mark, Starzinski-Powitz and Wixler2002; Petropoulos and Skerjanc, Reference Petropoulos and Skerjanc2002; Shi et al., Reference Shi, Bourdelas, Umbhauer and Boucaut2002). Hence the inhibition of GSK3β facilitates cell proliferation, muscle differentiation (via β-catenin) and, subsequently, hypertrophy.

Contractile activity and integrins

Muscle hypertrophy can be an adaptive response to load-bearing exercise, which stimulates the local expression of IGF-1, a potent inducer of muscle hyperplasia and hypertrophy. There is also growing evidence that key signalling intermediates in the Erk-MAPK and PI3K-Akt1 pathways can be activated independent of growth factor stimulation (Figures 2 and 3). Integrins are heterodimeric (αβ) transmembrane adhesion proteins that link the actin cytoskeleton to the extracellular matrix, and participate in a vast array of signalling events including mechanical sensing, cell growth and cell migration. Stimulation of β3 integrin by a synthetic Arg-Gly-Asp ligand in feline cardiomycytes led to activation of PI3K, mTOR, p70S6K and MEK/Erk (Balasubramanian and Kuppuswamy, Reference Balasubramanian and Kuppuswamy2003). It appears that in rats exposed to certain treadmill running regimes, and in certain isolated muscles subjected to passive stretch, Akt1 can be independently activated (Sakamoto et al., Reference Sakamoto, Aschenbach, Hirshman and Goodyear2003). mTOR, but not Akt1, was also found to be activated in rat hindlimb muscles subjected to 6 hours of high frequency electrical stimulation (Parkington et al., Reference Parkington, Siebert, LeBrasseur and Fielding2003). Hence exercise or electrical stimulation can variously activate the Erk-MAPK and PI3K-Akt1 pathways as well as the p38 MAPK cascade (see next section) in the regulation of muscle development.

p38 MAP kinase pathway of cell arrest and atrophy

p38 MAP kinase represents another MAPK signal transduction pathway that is activated by cellular stress (e.g. heat shock, oxidative stress and certain cytokines) as well as insulin (Conejo et al., Reference Conejo, Valverde, Benito and Lorenzo2001; Lee et al., Reference Lee, Hong, Kwon, Kim, Ki, Kang, Lee, Ha and Kim2002). The p38 MAPK pathway is intriguing in that it appears to mediate a range of cellular processes in muscle besides fibre size (Figure 4). There are four members of the p38 MAP kinase family: p38 MAPK-α, -β, -γ and -δ. The dual specificity kinases, MKK3 and MKK6, are involved in p38 MAP kinase activation, with MKK3 activating p38 MAPKα and β, and MKK6 activating all four p38 MAPK isoforms (Lee et al., Reference Lee, Hong, Kwon, Kim, Ki, Kang, Lee, Ha and Kim2002) (Figure 4). A variety of receptor families (including interleukin-1 [IL-1], tumour necrosis factor α [TNFα] and TGFβ) signal through this cascade. Signalling of p38 MAPK has been reported to be essential for terminal differentiation (myoblasts fusion into myotubes), by enhancing the expression of MyoD, MEF2A, MEF2C, sarcomeric muscle genes and cdk inhibitor p21 (Cabane et al., Reference Cabane, Englaro, Yeow, Ragno and Derijard2003; Wu et al., Reference Wu, Woodring, Bhakta, Tamura, Wen, Feramisco, Karin, Wang and Puri2000c). p38 MAPK has the added role of inhibiting the Erk-MAPK pathway thereby inducing cell cycle arrest during muscle differentiation (Lee et al., Reference Lee, Hong, Kwon, Kim, Ki, Kang, Lee, Ha and Kim2002). It is interesting to note that the Erk-MAPK pathway is also inhibited by activated Akt1 through the prevention of Raf and MEK phosphorylation (Rommel et al., Reference Rommel, Clarke, Zimmermann, Nuñez, Rossman, Reid, Moelling, Yancopoulos and Glass1999; Zimmermann and Moelling, Reference Zimmermann and Moelling1999)(Figure 3). TNF-α stimulated p38 MAPK signalling up-regulates MAFbx mRNA expression, which codes for a major E3 ubiquitin ligase responsible for muscle atrophy (Li et al., Reference Li, Chen, John, Moylan, Jin, Mann and Reid2005) (see section on markers of atrophy and hypertrophy). Exercise induced p38 MAPK signalling has been shown to stimulate PGC-1α expression which leads to mitochondrial biogenesis hence promoting oxidative capacity (Akimoto et al., Reference Akimoto, Pohnert, Li, Zhang, Gumbs, Rosenberg, Williams and Yan2005). Furthermore, we recently found that the p38 MAPK pathway performs an important role in the activation of the fast oxidative-glycolytic MyHC 2x promoter in skeletal muscle (J.D. Meissner et al., unpublished data). Hence, depending on the nature of p38 MAPK stimulation, different phenotypic outcomes, such as cell arrest, atrophy, terminal differentiation or raised oxidative capacity, can result from this pathway (Figure 4).

Synthetic β2 adrenergic agonists

β2 adrenergic receptor agonists (β2-agonists), such as clenbuterol, cimaterol and fenoterol, are potent agents for muscle hypertrophy as well as fibre type switch from slow/I to fast fibres (Ryall et al., Reference Ryall, Gregorevic, Plant, Sillence and Lynch2002). They are also important regulators of T-cells development in the thymus (Blanco et al., Reference Blanco, Artacho-Perula, Flores-Acuna, Moyano and Monterde2003), and were primarily developed for use as smooth muscle bronchodilators (Ryall et al., Reference Ryall, Gregorevic, Plant, Sillence and Lynch2002). A feature of β adrenergic receptors signalling, through the binding of catecholamines, is triglyceride hydrolysis, a process that has been harnessed in animal production to produce leaner carcasses. Ractopamine (‘Paylean’, Elanco), a β1- and β2-agonist, is a commercial compound sold in the USA as a promoter of lean growing pigs (Mills et al., Reference Mills, Spurlock and Smith2003). β2-adrenergic receptor but not β1 adrenergic receptor is responsible for mediating the hypertrophic effect of clenbuterol (Hinkle et al., Reference Hinkle, Hodge, Cody, Sheldon, Kobilka and Isfort2002). The growth promoting efficacy of β2-agonists appears to show animal species variation; ruminants display the greatest and broiler chickens the least response, with pigs occupying an intermediate position (Mersmann, Reference Mersmann1998). Although the effects of β-agonists on meat quality are somewhat equivocal, the overall picture, in particular in the pig, is that many β-agonists decrease intramuscular fat and increase shear force or toughness (Dunshea et al., Reference Dunshea, D'Souza, Pethick, Harper and Warner2005).

The signalling events of β2-agonists leading to muscle hypertrophy is poorly understood. There is evidence to suggest that clenbuterol induces local muscle production of IGF-1, which mediates hypertrophy (Awede et al., Reference Awede, Thissen and Lebacq2002). More recent work, however, could not detect sustained local production of IGF-1 in rat muscles treated with clenbuterol but a reduction in the expression of components of the ubiquitin-proteasome pathway was found (Yimlamai et al., Reference Yimlamai, Dodd, Borst and Park2005). The use of clenbuterol in meat production or in the treatment of muscle wasting conditions, however, has been impaired by reports of possible side effects. Pigs treated with anabolic doses of clenbuterol showed immunosuppressive effects, with raised T-cell apoptotic index (Blanco et al., Reference Blanco, Artacho-Perula, Flores-Acuna, Moyano and Monterde2003) and testicular degeneration (Blanco et al., Reference Blanco, Flores-Acuna, Roldan-Villalobos and Monterde2002). In the pig, the hypertrophic effect of clenbuterol on muscle was short-lived after drug withdrawal, which could render its use ineffective in the growth promotion of food animals (Sillence et al., Reference Sillence, Munn and Campbell2002). In rats, clenbuterol was found to cause significant cardiac and skeletal muscle necrosis (Burniston et al., Reference Burniston, Ng, Clark, Colyer, Tan and Goldspink2002).

Anabolic steroids

Testosterone is an important male hormone whose effects on body composition are to increase muscle mass and reduce fat, effects that are not dissimilar to β2-agonists (Kutscher et al., Reference Kutscher, Luna and Perry2002). Testosterone administration is associated with hypertrophy of type 1 and 2 fibres, and increases in satellite cell number in humans and some animals (Sinha-Hikim et al., Reference Sinha-Hikim, Roth, Lee and Bhasin2003; Chen et al., Reference Chen, Zajac and MacLean2005). In porcine satellite cells, however, the effect of testosterone on proliferation rate is equivocal (Doumit et al., Reference Doumit, Cook and Merkel1996). Anabolic androgenic steroids (stanozolol and nandrolone), made notorious by their misuse in sports, are structural synthetic derivatives of testosterone, designed to maximise anabolic and reduce androgenic (male sexual) effects. Outside the European Union, where anabolic steroids are used in beef production, meat quality can be adversely affected by a small but significant rise in shear force (Dunshea et al., Reference Dunshea, D'Souza, Pethick, Harper and Warner2005). Little is known about the anabolic steroid-induced signalling mechanisms that regulate muscle mass. One of the primary anabolic effects of androgens may be their ability to stimulate localised IGF-1 production in skeletal muscle (Chen et al., Reference Chen, Zajac and MacLean2005).

Ski

Ski is a nuclear proto-oncogene that is essential for embryonic development. It is expressed in most post-natal tissues. Like IGF-1, it exhibits dual functions in promoting cell proliferation and muscle differentiation. In addition to transforming chicken embryo fibroblasts, Ski can induce cells derived from quail embryonic body wall to undergo myogenesis. Lines of transgenic mice carrying chicken Ski cDNAs under the control of murine sarcoma virus (MSV) long terminal repeat (LTR) preferentially express high levels of Ski mRNA and protein in skeletal muscle, even though MSV-LTR is usually active in other tissues (Sutrave et al., Reference Sutrave, Kelly and Hughes1990 and Reference Sutrave, Leferovich, Kelly and Hughes2000). This near exclusive skeletal muscle expression of Ski is associated with hypertrophy of fast glycolytic 2b fibres, without increase in fibre number or nuclear number (Sutrave et al., Reference Sutrave, Kelly and Hughes1990). The hypertrophic effect of Ski on 2b fibres is associated with reduced protein degradation rates (Costelli et al., Reference Costelli, Carbo, Busquets, Lopez-Soriano, Baccino and Argiles2003). It could be that Ski over-expression is deleterious to cells, which would account for its restricted distribution of expression in MSV-LTR-Ski transgenic mice, and an absence of expression in skeletal-α-actin promoter-driven Ski transgenic mice (Sutrave et al., Reference Sutrave, Leferovich, Kelly and Hughes2000). Indeed, satellite cells isolated from muscles of MSV-LTR-Ski transgenic mice showed accelerated deterioration in termination differentiation with increasing age (Charge et al., Reference Charge, Brack and Hughes2002).

Unlike other oncogenes, the over-expression of the wild type c-Ski is sufficient to cause transformation (Prunier et al., Reference Prunier, Pessah, Ferrand, Seo, Howe and Atfi2003). The oncoproteins from Ski and the related SnoN (ski-related novel gene) are able to interact with a variety of transcription regulatory complexes, including histone deacetylase complexes (HDACs) and tumour suppressors (Ueki and Hayman, Reference Ueki and Hayman2003). The TGF-β signalling pathway has been identified as a key interacting site of Ski. The regulation of cell growth and differentiation by TGF-β is mediated by the Smad proteins, which are important tumour suppressors. TGF-β signalling is initiated when the bound ligand induces the formation of a heteromeric complex comprising type I and type II serine/threonine receptors. Type II receptor transphosphorylates type I receptor, which in turn phosphorylates Smad2 and Smad3. Activated Smad2 and Smad3 form heterodimers with Smad4 and translocate into the nucleus where they interact with a host of complexes to bring about transcriptional activation or repression of specific genes (Figure 4). Ski was recently found to directly interact with Smad2, Smad3 and Smad4, and to block the phosphorylation of Smad2 and Smad3 by activated TGF-β type I receptor (Prunier et al., Reference Prunier, Pessah, Ferrand, Seo, Howe and Atfi2003; Ueki and Hayman, Reference Ueki and Hayman2003). Therefore, a mechanism of the transforming ability of Ski (and SnoN) is the repression of Smad function, whose inactivation prevents TGF-β-induced cell cycle arrest. The up-regulation of Ski in proliferating satellite cells points to its possible role in mediating cell proliferation in the regeneration of damaged fibres (Soeta et al., Reference Soeta, Suzuki, Suzuki, Naito, Tachi and Tojo2001). The mechanisms behind the effects of Ski on muscle differentiation and hypertrophy are unknown. One speculation is that Ski inhibits the signalling of myostatin, a member of the TGF-β superfamily, thereby enhancing cell proliferation and differentiation (Costelli et al., Reference Costelli, Carbo, Busquets, Lopez-Soriano, Baccino and Argiles2003) (Figure 4). Ski might be developed as a candidate marker for hypertrophic growth in marker-assisted selection.

Fibre types: coordinated isoform-specific expression

The plasticity of muscle fibres is not confined to its ability to undergo changes in fibre size. Physiological and biochemical properties can show wide variations between individual fibres, and such variations are further subjected to modulations by external stimuli. Traditionally, classification of muscle fibre types is based on differences in a number of biochemical parameters between fibres (Gil et al., Reference Gil, Lopez-Albors, Vazquez, Latorre, Ramirez-Zarzosa and Moreno2001; Zierath and Hawley, Reference Zierath and Hawley2006). Succinate dehydrogenase (SDH) histochemical staining, for example, is able to differentiate fibres into two or three different types, based on the relative amount of the enzyme present in each fibre. As SDH is an integral component of the citric acid cycle, strongly positive fibres are classified as oxidative fibres. Another commonly cited histochemical staining method depends on the overall myosin adenosine triphosphatase (ATPase) activity in each fibre (Brooke and Kaiser, Reference Brooke and Kaiser1970; Bancroft and Gamble, Reference Bancroft and Gamble2002). Myosin ATPase activity originates from myosin heavy chain (MyHC), the principal sacromeric protein component of the thick myofilament. Differential myosin ATPase staining is due to differences in susceptibility to pH between different MyHC isoforms. Since myosin ATPase is highly sensitive to pH change, this staining method is intrinsically prone to variability in results. Histochemical methods, such as SDH and myosin ATPase stainings, are invaluable in describing the biochemical profile of individual fibres. However, they are less reliable in the objective determination of fibre types. Different histochemical stains often provide slightly different classification outcome of individual fibres, which in the past had made findings of association studies between fibre types and meat quality traits variable and even contradictory (Essén-Gustavsson, Reference Essén-Gustavsson, Puolanne and Demeyer1993; Klont et al., Reference Klont, Brocks and Eikelenboom1998; Lefaucheur et al., Reference Lefaucheur, Milan, Eolan and Le Callennec2004).

Definition of fibre types. A recent major advance in farm animal muscle research has been the development of an objective approach to muscle fibre typing based on the identity of the primary MyHC isoform expressed in each fibre (Chang et al., Reference Chang, Fernandes and Goldspink1993 and Reference Chang, Fernandes and Dauncey1995; Chang and Fernandes, Reference Chang and Fernandes1997). MyHCs are encoded by a highly conserved multigene family, of which eight isoforms are known in mammals (2a, 2x, 2b, embryonic, perinatal, slow/β, extraocular and α), each with its own myosin ATPase activity and each encoded by a distinct gene (Weiss and Leinwand, Reference Weiss and Leinwand1996). In pre-natal mammalian muscles, the embryonic, perinatal and slow/β/type I MyHC isoforms represent the three dominant skeletal muscle fibre types in the developing foetus. Shortly after birth, the post-natal MyHC isoforms (2a, 2x and 2b) replace the expression of embryonic and perinatal MyHC genes. Thus, in post-natal muscles of pigs, dogs and rodents, there are four major fibre types (Figure 5) characterised by the expression of the slow/β/type I, 2a, 2x and 2b MyHC gene isoforms (Schiaffino and Reggiani, Reference Schiaffino and Reggiani1996; Wu et al., Reference Wu, Crumley and Caiozzo2000b). In cattle and horses, MyHC 2b fibres are effectively absent (Chikuni et al., Reference Chikuni, Muroya and Nakajima2004; Maccatrozzo et al., Reference Maccatrozzo, Patruno, Toniolo, Reggiani and Mascarello2004). Based on the MyHC approach, post-natal muscle fibres in animals can be resolved by immunocytochemistry or in situ hybridisation into three or four major types, depending on animal species. Metabolic, biochemical and biophysical characteristics, such as oxidative and glycolytic capacities, fibre size, colour, and glycogen and lipid contents, have been found to vary between MyHC fibre types (Schiaffino and Reggiani, Reference Schiaffino and Reggiani1996; Klont et al., Reference Klont, Brocks and Eikelenboom1998; Karlsson et al., Reference Karlsson, Klont and Fernandez1999) (Table 1). The slow/β and fast 2b fibres, also known as slow oxidative (red) and fast glycolytic (white) respectively, represent two extreme metabolic profiles. Slow MyHC fibres are characterised by slow isoform contractile proteins, high levels of myoglobin, high volumes of mitochondria, high oxidative capacity, high lipid contents and high capillary density. Favourable meat traits such as colour and, in the pig in particular, tenderness have been found to closely associate with the greater abundance of red or highly oxidative fibres (Klont et al., Reference Klont, Brocks and Eikelenboom1998; Karlsson et al., Reference Karlsson, Klont and Fernandez1999; Chang et al., Reference Chang, Da Costa, Blackley, Southwood, Evans, Plastow, Wood and Richardson2003; Maltin et al., Reference Maltin, Balcerzak, Tilley and Delday2003; Wood et al., Reference Wood, Nute, Richardson, Whittington, Southwood, Plastow, Mansbridge, da Costa and Chang2004). There is a general perception that leaner meat, especially pork, contains reduced intramuscular fat, resulting in increased toughness and reduced succulence (Dunshea et al., Reference Dunshea, D'Souza, Pethick, Harper and Warner2005). Red muscles possess higher lipid concentration (intra- and inter-fibre fat) which is associated with more tender / juicy meat (Hocquette et al., Reference Hocquette, Ortigues-Marty, Pethick, Herpin and Fernandez1998;Wood et al., Reference Wood, Enser, Fisher, Nute, Richardson and Sheard1999 and Reference Wood, Richardson, Nute, Fisher, Campo, Kasapidou, Sheard and Enser2003).

Figure 5 Typing of porcine muscle fibres. Serial sections of the psoas of a 22-week-old pig to illustrate the combined use of histochemistry, immunocytochemistry and in situ hybridisation to identify the four major post-natal MyHC fibre types. Succinate dehydrogenase (SDH) staining (a), sudan black staining (b), NOQ7.5.4D (slow) monoclonal antibody binding (c), and in situ hybridisations with MyHC isoform-specific fast 2a probe (d), fast 2b (e) and fast 2x probe (f) were performed. SDH and Sudan Black positive fibres closely mirrored each other and corresponded mainly to slow and fast 2a fibres. Note the presence of inter-fibre adipose tissue on panel b.

Table 1 Metabolic profiles of MyHC fibre types

Pale, soft exudative (PSE) meat quality, independent of ryanodine receptor mutation. The abundance of MyHC 2b fibres in a normal pig is a contributory factor to PSE. ( ) indicates possibly variable levels or data not fully established in the pig.

By contrast, fast MyHC 2b fibres are the largest of the four fibre types with fast isoform contractile proteins, low amounts of myoglobin and mitochondria, high glycolytic capacity (high glycogen store), low lipid contents and low capillary density. The fast MyHC 2a and 2x fibres are intermediate fast oxidative-glycolytic fibres. Fast 2a fibres are more closely related to slow/I fibres, and fast 2x are more similar to fast 2b fibres (Table 1). Fast glycolytic fibres, in particular 2b fibres, are major contributors of hypertrophic growth and of rapid fall in muscle pH post mortem, associated with the formation of PSE pork. Hence in the modern pig, hypertrophic growth potential (meat quantity) from an abundance of MyHC2b and 2x fibres comes at a cost to meat quality (Chang et al., Reference Chang, Da Costa, Blackley, Southwood, Evans, Plastow, Wood and Richardson2003). On the other hand, the absence of MyHC 2b fibres in cattle is a likely explanation for the lack of a PSE problem in beef. Some beef studies have linked an abundance of fast-twitch fibres (defined by histochemical detection) to improved tenderness. The interpretation of fast-twitch fibres should be made with care as it is presently clear that in the absence of MyHC 2b fibres there are strictly speaking no fast-glycolytic fibres in bovine muscle (Geay et al., Reference Geay, Bauchart, Hocquette and Culioli2001; Maltin et al., Reference Maltin, Balcerzak, Tilley and Delday2003). Consequently, in a technical sense, bovine muscle may have higher oxidative potential than its porcine counterpart.

In addition to temporal regulation, each MyHC isoform is subjected to specific spatial control, such that the number and distribution of fibres expressing each isoform often vary between anatomical muscles, e.g. soleus and longissimus thoracis et lumborum (also referred by some as longissimus dorsi). Fibre type composition varies between muscles according to their functional adaptation. Postural muscles are under continual use and comprise a high proportion of oxidative fibres. Muscles that are periodically used for intensive activities like sprinting possess large numbers of fast fibres. A further degree of fibre type heterogeneity is the presence of a small number of hybrid MyHC fibres, usually presented as a mixture of two MyHC isoforms (slow/2a, 2a/2x, or 2x/2b) (Sant'ana Pereira et al., Reference Sant'ana Pereira, Wessels, Nijtmans, Moorman and Sargeant1995; Pette and Staron, Reference Pette and Staron2000). Thus fibre population in muscle is a continuum of pure and hybrid fibres that can be altered in the fast-to-slow or slow-to-fast direction under appropriate stimulatory conditions (Schiaffino and Reggiani, Reference Schiaffino and Reggiani1994).

Distinctive biochemical and biophysical differences between fibre types point to a coordinated programme of fibre-type or isoform specific expression (Hallauer and Hastings, Reference Hallauer and Hastings2002). However, compared with our knowledge of muscle hypertrophy, much less is known about the molecular mediators of fibre type specific expression. Coordinated fibre type specific expression requires the orchestrated regulation of a large number of gene isoforms consistent with the fibre phenotype. Gene family and differential splicing are features of muscle genes (Schiaffino and Reggiani, Reference Schiaffino and Reggiani1996). Understanding the complexities of fibre type specific expression, requires insights into the signalling pathways that coordinate the temporal and spatial distribution of expression of subsets of muscle gene isoforms. The emerging picture of fibre type specific regulation is that it is governed by multiple signalling pathways and factors rather than a single pathway (Spangenburg and Booth, Reference Spangenburg and Booth2003). These pathways and factors almost inevitably have additional cellular functions, such as proliferation, differentiation and hypertrophy, in addition to modulating fibre type specific expression.

IGF-1, β2-agonists and myostatin-null signal fast fibre phenotype

As highlighted earlier, the major phenotypic effects of IGF-1 and -2 are cell proliferation, muscle differentiation and hypertrophy. Such properties of IGF-1 have been shown to be beneficial to ageing, atrophic, and dystrophic muscles (Barton-Davis et al., Reference Barton-Davis, Shoturma, Musaro, Rosenthal and Sweeney1998; Lynch et al., Reference Lynch, Cuffe, Plant and Gregorevic2001; Musarò et al., Reference Musarò, McCullagh, Paul, Houghton, Dobrowolny, Molinaro, Barton, Sweeney and Rosenthal2001). IGF-1 has the additional effect of converting fibres to a fast glycolytic phenotype, as evident by raised expression of glycolytic enzymes in IGF-1-transfected C2C12 myotubes (Semsarian et al., Reference Semsarian, Wu, Ju, Marcinec, Yeoh, Allen, Harvey and Graham1999), by modest rise in fast 2b fibres in transgenic mice carrying muscle IGF-1 isoform driven by a rat myosin light chain (MLC)-1/3 promoter (Musarò et al., Reference Musarò, McCullagh, Paul, Houghton, Dobrowolny, Molinaro, Barton, Sweeney and Rosenthal2001) and by raised type 2a and 2b fibres at the expense of slow fibres in IGF-1-treated dystrophic mice (Lynch et al., Reference Lynch, Cuffe, Plant and Gregorevic2001). As previously indicated, clenbuterol (a β2-agonist) treated rats (Ryall et al., Reference Ryall, Gregorevic, Plant, Sillence and Lynch2002) as well as myostatin-null mice (Steelman et al., Reference Steelman, Recknor, Nettleton and Reecy2006) not only showed muscle hypertrophy but also slow-to-fast fibre conversion. The mechanism responsible for the slow-to-fast fibre type switch, however, remains elusive. In general, in the pig and ruminants, it appears that growth hormone (indirectly IGF-1) and β-agonists may negatively affect meat quality, as shown by increased shear force or sensory perception (Dunshea et al., Reference Dunshea, D'Souza, Pethick, Harper and Warner2005). Such effects may be connected to increased fast glycolytic (MyHC 2b or 2x) fibres but because fibre typing in meat quality association studies had been almost exclusively based on histochemical methods, the changes in MyHC 2b/2x fibres in the pig or MyHC 2x fibres in cattle have not been determined.

Calcineurin signals oxidative fibre type

Calcineurin (protein phosphatase 2B/PP2B) is an enzyme complex that comprises calcineurin A (CnA) catalytic subunit, calcineurin B (CnB) regulatory subunit and calcium-binding protein calmodulin (Schulz and Yutzey, Reference Schulz and Yutzey2004). It is a calcium dependent serine-threonine phosphatase that is widely distributed throughout the body. Calcineurin has been implicated in a wide variety of biological processes, including T-lymphocyte activation, vascular, neuronal and cardiac development and growth, and, more recently, skeletal muscle development (Crabtree, Reference Crabtree2001; Bueno et al., Reference Bueno, Van Rooij, Molkentin, Doevendans and De Windt2002; Horsley and Pavlath, Reference Horsley and Pavlath2002). In cardiac muscle, calcineurin signalling is necessary for cardiomyocyte maturation, heart chamber formation and cardiac hypertrophy (Schulz and Yutzey, Reference Schulz and Yutzey2004). In skeletal muscle, calcineurin appears to be needed in a number of key developmental processes, namely enhanced muscle cell differentiation, and in the fibre type context, conversion to slow (oxidative) muscle phenotype (Musarò et al., Reference Musarò, McCullagh, Naya, Olson and Rosenthal1999; Semsarian et al., Reference Semsarian, Wu, Ju, Marcinec, Yeoh, Allen, Harvey and Graham1999; Bigard et al., Reference Bigard, Sanchez, Zoll, Mateo, Rousseau, Veksler and Ventura-Clapier2000; Delling et al., Reference Delling, Tureckova, Lim, De Windt, Rotwein and Molkentin2000). It has been reported that activated calcineurin mediates the hypertrophic effect of IGF-1 (Musarò et al., Reference Musarò, McCullagh, Naya, Olson and Rosenthal1999; Semsarian et al., Reference Semsarian, Wu, Ju, Marcinec, Yeoh, Allen, Harvey and Graham1999). However, there is compelling evidence, including transgenic and knock-out data, to show that calcineurin has no effect on muscle hypertrophy but that the hypertrophic effect of IGF-1 is mediated by the PI3K pathway as detailed earlier (Naya et al., Reference Naya, Mercer, Shelton, Richardson, Williams and Olson2000; Bodine et al., Reference Bodine, Stitt, Gonzalez, Kline, Stover, Bauerlein, Zlotchenko, Scrimgeour, Lawrence, Glass and Yancopoulos2001b; Rommel et al., Reference Rommel, Bodine, Clarke, Rossman, Nunez, Stitt, Yancopoulos and Glass2001; Pallafacchina et al., Reference Pallafacchina, Calabria, Serrano, Kalhovde and Schiaffino2002). Calcineurin is activated by raised intracellular calcium, triggered by extracellular signals like nervous impulses, or hormonal input, such as IGF-1 stimulation. To date, few substrates, namely NFATs, MEF2s and PGC-1α, of calcineurin action in skeletal muscle are known.

NFATs. The best characterised calcineurin substrates are members of the phosphorylated NFAT (nuclear factor of activated T-cells) family of transcription factors. Five NFAT genes, each with a distinct cellular role, have been identified: NFATc1 (NFAT2/NFATc), NFATc2 (NFATp/NFAT1), NFATc3 (NFAT4/NFATx), NFATc4 (NFAT3), and NFAT5 (Delling et al., Reference Delling, Tureckova, Lim, De Windt, Rotwein and Molkentin2000; Horsley and Pavlath, Reference Horsley and Pavlath2002). Several NFAT isoforms are expressed in skeletal muscle, each of which undergoes activation at specific stages of myogenesis. For instance, NFATc2 is activated only in new myotubes and plays a crucial role via IL-4 in mediating myoblast fusion (Horsley et al., Reference Horsley, Jansen, Mills and Pavlath2003). Dephosphorylation of NFATs unmasks their nuclear localisation signal, resulting in nuclear translocation, where they bind to NFAT-binding sites on their own or as co-operative complexes with other factors, such as calcineurin, AP1, MEF2 and GATA2/4, to activate gene transcription (Musarò et al., Reference Musarò, McCullagh, Naya, Olson and Rosenthal1999; Sugiura et al., Reference Sugiura, Sio, Shuntoh and Kuno2001; Schulz and Yutzey, Reference Schulz and Yutzey2004). Cultured muscle fibres showed NFAT nuclear translocation only when electrically stimulated at a slow muscle pattern, which results in high intracellular calcium (100 to 300 nM) but not at a fast muscle pattern that results in low intracellular calcium ( < 50 nM) (Liu et al., Reference Liu, Cseresnyés, Randall and Schneider2001; Fraysse et al., Reference Fraysse, Desaphy, Pierno, De Luca, Liantonio, Mitolo and Camerino2003). Nuclear NFATs are deactivated and re-located back to the cytoplasm through phosphorylation by several protein kinases, including GSK-3β (Figure 3), protein kinase A, p38 MAP kinase and casein kinase (Schulz and Yutzey, Reference Schulz and Yutzey2004). In T-cells, simultaneous activation of AP-1 (Jun/Fos) and NFAT is absolutely essential for cytokine transcriptional induction (Rao et al., Reference Rao, Luo and Hogan1997). In skeletal muscle, however, only GATA2 has been shown to be a co-operative binding partner of NFATs (Paul and Rosenthal, Reference Paul and Rosenthal2002).

MEF2s. Members of the MEF2 (myocyte enhancer-binding factor 2) family belong to a class of transcription factors that are responsible for the activation of many muscle-specific genes with a conserved A/T-rich cis-acting regulatory element (Naya and Olson, Reference Naya and Olson1999) and play a crucial role in p38 MAP kinase-mediated terminal muscle differentiation (discussed above). Like NFATs, MEF2 proteins form co-operative complexes with other factors, like members of the MyoD family, to regulate transcription. Calcineurin up-regulates the transcription of MEF2 genes (Wu et al., Reference Wu, Naya, McKinsey, Mercer, Shelton, Chin, Simard, Michel, Bassel-Duby, Olson and Williams2000a). It can also dephosphorylate MEF2 directly and enhance its transactivational activity (Dunn et al., Reference Dunn, Simard, Bassel-Duby, Williams and Michel2001; Wu et al., Reference Wu, Rothermel, Kanatous, Rosenberg, Naya, Shelton, Hutcheson, DiMaio, Olson, Bassel-Duby and Williams2001). Hence members of the MEF2 family are transcriptional targets and protein substrates of calcineurin.

PGC-1α. Peroxisome proliferator-activated receptor-γ co-activator-1 (PPARγ co-activator 1α /PGC-1α), a ubiquitous transcriptional co-factor for nuclear receptors, is a potent inducer of mitochondrial biogenesis (Lin et al., Reference Lin, Wu, Tarr, Zhang, Wu, Boss, Michael, Puigserver, Isotani, Olson, Lowell, Bassel-Duby and Spiegelman2002). It was recently shown to be able to convert fast fibres to the slow phenotype when over-expressed in transgenic mice. Part of this effect is thought to be mediated through the action of calcineurin on PGC-1α as a substrate (see PPARs below) (Lin et al., Reference Lin, Wu, Tarr, Zhang, Wu, Boss, Michael, Puigserver, Isotani, Olson, Lowell, Bassel-Duby and Spiegelman2002). As a co-factor, PGC-1α is likely to exert its effects indirectly by modulating the expression of a specific group of downstream genes. Although all three substrates (NFATs, MEF2s and PGC-1α) are important mediators of calcineurin activation, none is exclusively expressed in skeletal muscle. Additional substrates not yet identified may well be involved in the signalling pathway. It is not known which regulatory target genes activated by NFAT, MEF-2 or PGC-1α are responsible for the phenotypic effects of calcineurin. Further work is needed to identify the genes regulated by the known and, possibly, other unidentified substrates of calcineurin.

Several endogenous calcineurin-specific inhibitors have been discovered. AKAP79, cain/cabin 1, and CBHP are ubiquitous factors and were found to inhibit NFAT function or translocation to the nucleus (Crabtree, Reference Crabtree2001). More recently, two additional endogenous calcineurin-specific inhibitors (DSCR1/MCIP1 and ZAKI-4/DSCR1L1/MCIP2), highly expressed in striated muscles and brain, were found (Yang et al., Reference Yang, Rothermel, Vega, Frey, McKinsey, Olson, Bassel-Duby and Williams2000; Rothermel et al., Reference Rothermel, McKinsey, Vega, Nicol, Mammen, Yang, Antos, Shelton, Bassel-Duby, Olson and Williams2001). DSCR1 expression is induced by calcineurin and hence forms a negative feedback loop to limit calcineurin activation (Yang et al., Reference Yang, Rothermel, Vega, Frey, McKinsey, Olson, Bassel-Duby and Williams2000). Its over-expression in transgenic mice prevented cardiac hypertrophy (Rothermel et al., Reference Rothermel, McKinsey, Vega, Nicol, Mammen, Yang, Antos, Shelton, Bassel-Duby, Olson and Williams2001; Van Rooij et al., Reference Van Rooij, Doevendans, Crijns, Heeneman, Lips, Van Bilsen, Williams, Olson, Bassel-Duby, Rothermel and De Windt2004). In skeletal muscle, the role of ZAKI-4 is particularly relevant because, unlike DSCR1, it is responsive to thyroid hormone stimulation (Cao et al., Reference Cao, Kambe, Miyazaki, Sarkar, Ohmori and Seo2002). Thyroid hormone is a well known pleiotropic endocrine regulator of metabolism that modulates the transcription of a large number of genes, leading to increased metabolic rate, protein breakdown and muscle loss (Clement et al., Reference Clement, Viguerie, Diehn, Alizadeh, Barbe, Thalamas, Storey, Brown, Barsh and Langin2002). The importance of thyroid hormone, however, on glucose and lipid homeostasis has not been addressed. In hyperthyroidism, muscle loss is accompanied by increase of fast fibres at the expense of slow fibres (Caiozzo et al., Reference Caiozzo, Baker, McCue and Baldwin1997). Triiodothyronine (T3) with clenbuterol greatly enhances the slow to fast fibre type switch (Awede et al., Reference Awede, Thissen and Lebacq2002). Such changes resemble the effects of chemical inhibitors of calcineurin, cyclosporine A or FK506, on muscle (Bueno et al., Reference Bueno, Van Rooij, Molkentin, Doevendans and De Windt2002). It is possible that the effects of thyroid hormone on muscle loss and fast phenotype conversion is mediated through the inhibition of calcinuerin by ZAKI-4.

There is little doubt that the calcineurin-NFAT pathway is critical in muscle phenotype determination, in particular its roles in muscle differentiation and slow fibre conversion. However, as in other pathways, calcineurin signalling is not an absolute effect in that it does not activate all slow genes in all muscles. Some fast muscle genes, like MyHC 2b and SERCA1, are up-regulated in activated calcineurin transfected C2C12 cells (Swoap et al., Reference Swoap, Hunter, Stevenson, Felton, Kansagra, Lang, Esser and Kandarian2000). Transgenic mice over-expressing the activated calcineurin in skeletal muscle showed substantial slow fibre switch but only in certain muscles (Naya et al., Reference Naya, Mercer, Shelton, Richardson, Williams and Olson2000).

PPARs and PGC-1α influence oxidative phenotype

Peroxisome proliferator-activated receptors (PPAR) α, γ and δ/β belong to an important family of nuclear hormone receptors (transcription factors) that regulate genes that are involved in lipid metabolism (Muoio et al., Reference Muoio, Way, Tanner, Winegar, Kliewer, Houmard, Kraus and Dohm2002). PPARα, initially identified as the mediator of a class of compounds that induces peroxisomal proliferation in rodent liver, is most abundantly expressed in striated muscle (Muoio et al., Reference Muoio, Way, Tanner, Winegar, Kliewer, Houmard, Kraus and Dohm2002). Its activation increases fatty acid β-oxidation in muscle and reduces body weight (Koh et al., Reference Koh, Kim, Park, Kim, Youn, Park, Youn and Lee2003). PPARγ is more highly expressed in adipose tissues (Koh et al., Reference Koh, Kim, Park, Kim, Youn, Park, Youn and Lee2003) and, in contrast to PPARα, stimulates adipocyte differentiation (Oberkofler et al., Reference Oberkofler, Esterbauer, Linnemayr, Strosberg, Krempler and Patsch2002; Yu et al., Reference Yu, Liu, Mersmann and Ding2006). PPARδ is highly expressed in adipose tissue, heart and skeletal muscle. Recently, it was shown that activated PPARδ increases fatty acid oxidation in adipose tissue and muscle (Wang et al., Reference Wang, Lee, Tiep, Yu, Ham, Kang and Evans2003) and that during starvation its expression is up-regulated in skeletal muscle (Holst et al., Reference Holst, Luquet, Nogueira, Kristiansen, Leverve and Grimaldi2003). PPARα, PPARγ and PPARδ bind to the same consensus sequence, suggesting that mechanisms must exist that regulate PPAR type specificity (Oberkofler et al., Reference Oberkofler, Esterbauer, Linnemayr, Strosberg, Krempler and Patsch2002).

PPARγ co-activator 1α (PGC-1α), cloned from a brown fat cDNA library for its interaction with PPARγ, is a potent transcriptional activator that interacts with several nuclear hormone receptors, which include PPARs, steroid hormone receptors (glucocorticoid, oestrogen and mineralocorticoid), retinoic X receptor, thyroid hormone receptor (Oberkofler et al., Reference Oberkofler, Esterbauer, Linnemayr, Strosberg, Krempler and Patsch2002), as well as MEF2 (Handschin et al., Reference Handschin, Rhee, Lin, Tarr and Spiegelman2003). PGC-1α stimulates mitochondrial function and number (biogenesis), and fatty acid oxidation in cardiac and skeletal muscles by the induction of mitochondrial and nuclear genes involved in energy production pathways, including PPARα and nuclear respiratory factor-1 (NRF-1) (Miura et al., Reference Miura, Kai, Ono and Ezaki2003). Therefore, the phenotypic effect of PGC-1α activation, in partnership with nuclear hormone receptors, including PPARs, in skeletal muscle is increased oxidative capacity.

Markers of atrophy and hypertrophy

Markers of atrophy and hypertrophy are relevant to farm animal production from the perspective of disease and welfare. Animals suffer from a wide range of infectious and non-infectious diseases that are often manifested as reduced weight gain or net weight loss. Markers that can readily measure muscle atrophy or hypertrophy could have veterinary applications.

Biological markers that can predict the functional state of muscle are useful indicators in the diagnosis and monitoring of muscle conditions. Clearly, muscles that have undergone gross hypertrophy or atrophy would be apparent during examination. Muscles in the transitional process of remodelling are more difficult to recognise. Thus the use predictive biological markers would be valuable in identifying the underlying changes that are taking place in muscle.

E3 ubiquitin ligases as atrophic markers. Muscle atrophy is a consequence of a variety of conditions: denervation, injury, limb immobilisation, inactivity, glucocorticoid treatment, infection, diabetes, renal failure, cancer and ageing. In muscle atrophy, muscle loss is often the result of raised protein degradation and turn-over rather than simply due to reduced protein synthesis. Thus, atrophy is as much an active degradation process as a passive one from reduced stimulation of the anabolic (hyperplastic and hypertrophic) pathways. Of the different protein degradation pathways (Costelli et al., Reference Costelli, Carbo, Busquets, Lopez-Soriano, Baccino and Argiles2003), the ATP-dependent ubiquitin-proteasome proteolytic system is thought to be of major importance in skeletal muscle (Dehoux et al., Reference Dehoux, Van Beneden, Fernandez-Celemin, Lause and Thissen2003; Glass, Reference Glass2003a). In this process, ubiquitin (Ub) is activated by an ubiquitin-activating enzyme (E1 family) and conjugated to a substrate protein by ubiquitin-conjugating enzyme (E2 family) in conjunction with ubiquitin protein ligase (E3 ligase family). E3 ligases, of which hundreds have been identified, confer substrate specificity (Glass, Reference Glass2003a). During muscle atrophy, many of the genes encoding the Ub-proteasome pathway are up-regulated (Lecker et al., Reference Lecker, Jagoe, Gilbert, Gomes, Baracos, Bailey, Price, Mitch and Goldberg2004), including E2 (such as E214K and UbcH2) and E3 (such as E3α / UBr1) genes (Li et al., Reference Li, Chen, Li and Reid2003; Lecker et al., Reference Lecker, Jagoe, Gilbert, Gomes, Baracos, Bailey, Price, Mitch and Goldberg2004). In particular, two E3 ubiquitin ligases, MuRF1 (muscle ring finger 1) and MAFbx (muscle atrophy F-box or atrogin-1) have been identified as mediators of muscle atrophy (Bodine et al., Reference Bodine, Latres, Baumhueter, Lai, Nunez, Clarke, Poueymirou, Panaro, Na, Dharmarajan, Pan, Valenzuela, DeChiara, Stitt, Yancopoulos and Glass2001a; Dehoux et al., Reference Dehoux, Van Beneden, Fernandez-Celemin, Lause and Thissen2003). MuRF1 and MAFbx are selectively induced in skeletal muscle and heart when subjected to a variety of atrophic conditions, such as denervation and glucocorticoid treatment. Their absence in null knock-out mice led to the preservation of muscle mass under atrophying conditions (Bodine et al., Reference Bodine, Latres, Baumhueter, Lai, Nunez, Clarke, Poueymirou, Panaro, Na, Dharmarajan, Pan, Valenzuela, DeChiara, Stitt, Yancopoulos and Glass2001a). TNF-α stimulated proteolysis may be mediated through the increased expression of MAFbx and MuRF1 (Dehoux et al., Reference Dehoux, Van Beneden, Fernandez-Celemin, Lause and Thissen2003; Glass Reference Glass2005). MuRF1 contains three domains: a RING-finger domain, required for ubiquitin ligase activity, a B-box of unclear function, and a coil-coil domain, which may be required for heterodimerisation with a related protein MuRF2. Activation, by cachectic factors such as TNFα, of the NF-κB transcription pathway induces skeletal muscle atrophy, which is mediated in part by NF-κB on the up-regulation of MuRF1 (Glass, Reference Glass2005). MuRF1 physically interacts with titin, which suggests a possible role for MuRF1 in titin turn-over (Centner et al., Reference Centner, Yano, Kimura, McElhinny, Pelin, Witt, Bang, Trombitas, Granzier and Gregorio2001). MAFbx contains an F-box domain, characteristic of SCF family of E3 ubiquitin ligases. F-box E3 ligases usually bind post-translationally modified substrates, such as phosphorylation, and may target proteins involved in cell signalling. We recently found that in porcine congenital splayleg, a condition characterised with severe muscle weakness and fibre atrophy, MAFbx is abnormally up-regulated in relation to normal littermates (Ooi et al., Reference Ooi, Da Costa, Edgar and Chang2006).

FOXO proteins, a subgroup of the Forkhead family of transcription factors, have been recently identified as mediators that link the signalling pathways of muscle hypertrophy and atrophy. FOXO transcription factors are important for the induction of cell quiescence and apoptosis. FOXO factors are targets of activated Akt (Figure 3). Direct and multiple phosphorylation by Akt on FOXO factors leads to their displacement from the nucleus to the cytoplasm and to the inhibition of their transcriptional activities (Burgering and Medema, Reference Burgering and Medema2004). Under atrophic conditions, FOXO1 (Stitt et al., Reference Stitt, Drujan, Clarke, Panaro, Timofeyva, Kline, Gonzalez, Yancopoulos and Glass2004) and FOXO3 (Sandri et al., Reference Sandri, Sandri, Gilbert, Skurk, Calabria, Picard, Walsh, Schiaffino, Lecker and Goldberg2004) have been shown to up-regulate the expression of MAFbx and MuRF1. Inhibition of FOXO activity that leads to the down-regulation of MAFbx and MuRF1 is a necessary anti-atrophy step mediated by the PI3K-Akt1 pathway (Sandri et al., Reference Sandri, Sandri, Gilbert, Skurk, Calabria, Picard, Walsh, Schiaffino, Lecker and Goldberg2004; Stitt et al., Reference Stitt, Drujan, Clarke, Panaro, Timofeyva, Kline, Gonzalez, Yancopoulos and Glass2004; Tesseraud et al., Reference Tesseraud, Metayer, Duchene, Bigot, Grizard and Dupont2007). This anti-atrophy effect can be triggered by insulin as ligand which has the dual roles of enhancing protein synthesis (Figure 3) and decreasing proteolysis via FOXO inhibition (Tesseraud et al., Reference Tesseraud, Metayer, Duchene, Bigot, Grizard and Dupont2007).

GATA-2 as a hypertrophic marker. Factors involved in signalling, muscle metabolism or sarcomeric structure that are associated with the process of muscle hypertrophy could potentially be used as hypertrophic markers. GATA-2 has been recently identified as a potentially useful candidate marker gene of muscle hypertrophy (Musarò et al., Reference Musarò, McCullagh, Naya, Olson and Rosenthal1999; Paul and Rosenthal, Reference Paul and Rosenthal2002). In mammals, the GATA zinc-finger transcription factor family comprises six gene members that can be divided into 2 sub-groups, based on structure and function (LaVoie, Reference LaVoie2003). GATA-1/-2/-3 members are often associated with haematopoiesis and neural development. GATA-4/-5/-6 members are commonly associated with organ development, including heart, gut, blood vessels and parts of the genito-urinary system. With the exception of GATA-5, gene ablation of the GATA family members results in embryonic lethality. GATA-2 is necessary for mast cell development, and maintenance and expansion of multipotential progenitors and haematopoietic stem cells (Fujiwara et al., Reference Fujiwara, Chang, Williams and Orkin2004). GATA-2 is usually absent or lowly expressed in skeletal muscle at all stages of development, but is induced in muscle cell culture and in vivo by IGF-1, during muscle regeneration after bupivacaine injection, and by exercise (Paul and Rosenthal, Reference Paul and Rosenthal2002; Sakuma et al., Reference Sakuma, Nishikawa, Nakao, Watanabe, Totsuka, Nakano, Sano and Yasuhara2003). GATA-2 co-precipitates with calcineurin and NFATc1 (see below), suggesting it mediates its effect on muscle gene expression in conjunction with components of the calcineurin signalling pathway as a protein complex (Musarò et al., Reference Musarò, McCullagh, Naya, Olson and Rosenthal1999; Sakuma et al., Reference Sakuma, Nishikawa, Nakao, Watanabe, Totsuka, Nakano, Sano and Yasuhara2003). There is no direct evidence to show that GATA-2 is a hypertrophic factor, but its expression is closely associated with factors (e.g. IGF-1) and conditions (e.g. exercise, regeneration) that are involved in the process of hypertrophy and regeneration.

Conclusion

The determination of muscle phenotype (fibre number, size and fibre type) is highly complex and coordinated that requires the integration of several major signalling cascades (e.g. Erk-MAPK, p38 MAPK, PI3K-Akt and calcineurin signalling pathways), intracellular factors (transcription factors, like NFATs, and co-factors, such as PGC-1α) and extracellular factors (e.g. ligands, like myostatin and IGFs, and nutrition) (Figure 6). Enhanced hypertrophic growth, as exemplified in the pig, for greater lean meat production is associated with reduced intramuscular fat and increased accumulation of fast glycolytic fibres, the most common being the MyHC 2x and 2b fibres (Chang et al., Reference Chang, Da Costa, Blackley, Southwood, Evans, Plastow, Wood and Richardson2003; Wood et al., Reference Wood, Nute, Richardson, Whittington, Southwood, Plastow, Mansbridge, da Costa and Chang2004). Highly glycolytic fibres in the pig, however, are not conducive to the conferment of good meat quality traits, such as colour and water-holding capacity. In future farm animal production, improvements on meat quality could offer considerable economic attraction. At a fundamental level, one need is to discover key targets that mediate slow or oxidative fibre type switching. The recent introduction of the use of exon-expression arrays and ChIP-on-chip tiling arrays (Affymetrix) in functional genomics is likely to greatly accelerate our understanding of key molecular and signalling details in fibre phenotype determination. Identified target genes could be exploited through marker-assisted selection or by pharmacological / nutritional manipulation.

Figure 6 Summary actions of signalling factors and pathways on fibre size, fibre number and fibre types. p38 MAPK activation can result in a range of phenotypic outcome depending on the nature of activation. The proposed inhibition of myostatin by Ski has not been formally demonstrated.

References

Adams, GR 2002. Autocrine and/or paracrine insulin-like growth factor-I activity in skeletal muscle. Clinical Orthopaedics and Related Research 403S, S188-S196.CrossRefGoogle Scholar
Akimoto, T, Pohnert, SC, Li, P, Zhang, M, Gumbs, C, Rosenberg, PB, Williams, RS and Yan, Z 2005. Exercise stimulates Pgc-1α transcription in skeletal muscle through activation of the p38 MAPK pathway. Journal of Biological Chemistry 280, 19587-19593.CrossRefGoogle ScholarPubMed
Artaza, JN, Bhasin, S, Magee, TR, Reisz-Porszasz, S, Shen, R, Groome, NP, Fareez, MM and Gonzalez-Cadavid, NF 2005. Myostatin inhibits myogenesis and promotes adipogenesis in C3H 10T(1/2) mesenchymal multipotent cells. Endocrinology 146, 3547-3557.CrossRefGoogle ScholarPubMed
Asakura, A 2003. Stem cells in adult skeletal muscle. Trends in Cardiovascular Medicine 13, 123-128.CrossRefGoogle ScholarPubMed
Awede, BL, Thissen, JP and Lebacq, J 2002. Role of IGF-I and IGFBPs in the changes of mass and phenotype induced in rat soleus muscle by clenbuterol. AJP - Endocrinology and Metabolism 282, E31-E37.CrossRefGoogle Scholar
Balasubramanian, S and Kuppuswamy, D 2003. RGD-containing peptides activate S6K1 through β3 integrin in adult cardiac muscle cells. Journal of Biological Chemistry 278, 42214-42224.Google Scholar
Bancroft, JD and Gamble, M 2002. Theory and practice of histological techniques. Churchill Livingston, London.Google Scholar
Barton-Davis, ER, Shoturma, DI, Musaro, A, Rosenthal, N and Sweeney, HL 1998. Viral mediated expression of insulin-like growth factor I blocks the aging-related loss of skeletal muscle function. Proceedings of the National Academy of Sciences USA 95, 15603-15607.CrossRefGoogle ScholarPubMed
Bass, J, Oldham, J, Sharma, M and Kambadur, R 1999. Growth factors controlling muscle development. Domestic Animal Endocrinology 17, 191-197.Google Scholar
Baxendale, S, Davison, C, Muxworthy, C, Wolff, C, Ingham, W and Roy, S 2004. The B-cell maturation factor Blimp-1 specifies vertebrate slow-twitch muscle fiber identity in response to Hedgehog signaling. Nature Genetics 36, 88-93.CrossRefGoogle ScholarPubMed
Beuzen, ND, Stear, MJ and Chang, KC 2000. Molecular markers and their use in animal breeding. Veterinary Journal 160, 42-52.CrossRefGoogle ScholarPubMed
Bigard, X, Sanchez, H, Zoll, J, Mateo, P, Rousseau, V, Veksler, V and Ventura-Clapier, R 2000. Calcineurin co-regulates contractile and metabolic components of slow muscle phenotype. Journal of Biological Chemistry 275, 19653-19660.CrossRefGoogle ScholarPubMed
Blanco, A, Artacho-Perula, E, Flores-Acuna, R, Moyano, R and Monterde, JG 2003. Quantitative changes in the normal and apoptotic thymocytes of pigs treated with anabolic doses of the β2 adrenergic agonist clenbuterol. Veterinary Immunology and Immunopathology 96, 111-115.CrossRefGoogle ScholarPubMed
Blanco, A, Flores-Acuna, F, Roldan-Villalobos, R and Monterde, JG 2002. Testicular damage from anabolic treatments with the β2-adrenergic agonist clenbuterol in pigs: a light and electron microscope study. Veterinary Journal 163, 292-298.Google Scholar
Bodine, SC, Latres, E, Baumhueter, S, Lai, VKM, Nunez, L, Clarke, BA, Poueymirou, WT, Panaro, FJ, Na, E, Dharmarajan, K, Pan, Z-Q, Valenzuela, DM, DeChiara, TM, Stitt, TN, Yancopoulos, GD and Glass, DJ 2001a. Identification of ubiquitin ligases required for skeletal muscle atrophy. Science 294, 1704-1708.Google Scholar
Bodine, SC, Stitt, TN, Gonzalez, M, Kline, WO, Stover, GL, Bauerlein, R, Zlotchenko, E, Scrimgeour, A, Lawrence, JC, Glass, DJ and Yancopoulos, GD 2001b. Akt/mTOR pathway is a crucial regulator of skeletal muscle hypertrophy and can prevent muscle atrophy in vivo. Nature Cell Biology 3, 1014-1019.Google Scholar
Bouley, J, Meunier, B, Chambon, C, De Smet, S, Hocquette, JF and Picard, B 2006. Proteomic analysis of bovine skeletal muscle hypertrophy. Proteomics 5, 490-500.CrossRefGoogle Scholar
Brooke, MH and Kaiser, KK 1970. Muscle fiber types: how many and what kind? Archives of Neurology 23, 369-379.CrossRefGoogle ScholarPubMed
Buckingham, M 2001. Skeletal muscle formation in vertebrates. Current Opinion in Genetics & Development 11, 440-448.CrossRefGoogle ScholarPubMed
Buckingham, M, Bajard, L, Chang, T, Daubas, P, Hadchouel, J, Meilhac, S, Montarras, D, Rocancourt, D and Relaix, F 2003. The formation of skeletal muscle: from somite to limb. Journal of Anatomy 202, 59-68.Google Scholar
Bueno, OF, Van Rooij, E, Molkentin, JD, Doevendans, PA and De Windt, LJ 2002. Calcium and hypertrophic heart disease: novel insights and remaining questions. Cardiovascular Research 53, 806-821.Google Scholar
Burgering, BMT and Medema, RH 2004. Decisions on life and death: FOXO forkhead transcription factors are in command when PKB/Akt is off duty. Journal of Leukocyte Biology 73, 689-701.CrossRefGoogle Scholar
Burniston, JG, Ng, Y, Clark, WA, Colyer, J, Tan, LB and Goldspink, DF 2002. Myotoxic effects of clenbuterol in the rat heart and soleus muscle. Journal of Applied Physiology 93, 1824-1832.Google Scholar
Byrne, KA, Wang, YH, Lehnert, SA, Harper, GS, McWilliam, SM, Bruce, HL and Reverter, A 2005. Gene expression profiling of muscle tissue in Brahman steers during nutritional restriction. Journal of Animal Science 83, 1-12.CrossRefGoogle ScholarPubMed
Cabane, C, Englaro, W, Yeow, K, Ragno, M and Derijard, B 2003. Regulation of C2C12 myogenic terminal differentiation by MKK3/p38α pathway. American Journal of Physiology -Cell Physiology 284, C658-C666.CrossRefGoogle ScholarPubMed
Caiozzo, VJ 2004. Plasticity of skeletal muscle phenotype: mechanical consequences. Muscle Nerve 26, 740-768.Google Scholar
Caiozzo, VJ, Baker, MJ, McCue, SA and Baldwin, KM 1997. Single-fiber and whole muscle analyses of MHC isoform plasticity: interaction between T3 and unloading. American Journal of Physiology -Cell Physiology 273, C944-C952.CrossRefGoogle ScholarPubMed
Cao, X, Kambe, F, Miyazaki, T, Sarkar, D, Ohmori, S and Seo, H 2002. Novel human ZAKI-4 isoforms: hormonal and tissue-specific regulation and function as calcineurin inhibitors. Biochemical Journal 367, 459-466.CrossRefGoogle ScholarPubMed
Casser-Malek, I, Hocquette, JF, Jurie, C, Listrat, A, Jailler, R, Bauchart, D, Briand, Y and Picard, B 2004. Muscle-specific metabolic, histochemical and biochemical responses to a nutritionally induced discontinuous growth path. Animal Science 79, 49-59.Google Scholar
Centner, T, Yano, J, Kimura, E, McElhinny, AS, Pelin, K, Witt, CC, Bang, ML, Trombitas, K, Granzier, H and Gregorio, CC 2001. Identification of muscle specific ring finger proteins as potential regulators of the titin kinase domain. Journal of Molecular Biology 306, 717-726.CrossRefGoogle ScholarPubMed
Chang, KC, Da Costa, N, Blackley, R, Southwood, O, Evans, G, Plastow, G, Wood, JD and Richardson, RI 2003. Relationships of myosin heavy chain fibre types to meat quality traits in traditional and modern pigs. Meat Science 64, 93-103.CrossRefGoogle ScholarPubMed
Chang, KC and Fernandes, K 1997. Developmental expression and 5′ end cDNA cloning of the porcine 2x and 2b myosin heavy chain genes. DNA and Cell Biology 16, 1429-1437.CrossRefGoogle ScholarPubMed
Chang, KC, Fernandes, K and Dauncey, MJ 1995. Molecular characterization of a developmentally regulated porcine skeletal myosin heavy chain gene and its 5′ regulatory region. Journal of Cell Science 108, 1779-1789.Google Scholar
Chang, KC, Fernandes, K and Goldspink, G 1993. In vivo expression and molecular characterization of the porcine slow-myosin heavy chain. Journal of Cell Science 106, 331-341.CrossRefGoogle ScholarPubMed
Charge, SBP, Brack, AS and Hughes, SM 2002. Aging-related satellite cell differentiation defect occurs prematurely after Ski-induced muscle hypertrophy. American Journal of Physiology -Cell Physiology 283, C1228-C1241.CrossRefGoogle ScholarPubMed
Chen, Y, Zajac, JD and MacLean, HE 2005. Androgen regulation of satellite cell function. Journal of Endocrinology 186, 21-31.Google Scholar
Chikuni, K, Muroya, S and Nakajima, I 2004. Absence of the functional myosin heavy chain 2b isoform in equine skeletal muscles. Zoological Science 21, 589-596.Google Scholar
Clement, K, Viguerie, N, Diehn, M, Alizadeh, A, Barbe, P, Thalamas, C, Storey, JD, Brown, PO, Barsh, GS and Langin, D 2002. In vivo regulation of human skeletal muscle gene expression by thyroid hormone. Genome Research 12, 281-291.Google Scholar
Conejo, R, Valverde, AM, Benito, M and Lorenzo, M 2001. Insulin produces myogenesis in C2C12 myoblasts by induction of NF-κB and downregulation of AP-1 activities. Journal of Cellular Physiology 186, 82-94.3.0.CO;2-R>CrossRefGoogle ScholarPubMed
Coolican, SA, Samuel, DS, Ewton, DZ and McWade, FJ 1997. The mitogenic and myogenic actions of insulin-like growth factors utilize distinct signaling pathways. Journal of Biological Chemistry 272, 6653-6662.Google Scholar
Costelli, P, Carbo, N, Busquets, S, Lopez-Soriano, FJ, Baccino, FM and Argiles, JM 2003. Reduced protein degradation rates and low expression of proteolytic systems support skeletal muscle hypertrophy in transgenic mice overexpressing the c-ski oncogene. Cancer Letters 200, 153-160.CrossRefGoogle ScholarPubMed
Crabtree, GR 2001. Calcium, calcineurin, and the control of transcription. Journal of Biological Chemistry 276, 2313-2316.Google Scholar
Da Costa, N, McGillivray, C, Bai, Q, Wood, JD, Evans, G and Chang, KC 2004. Energy and protein restriction induces molecular changes in young porcine skeletal muscles. Journal of Nutrition 134, 2191-2199.CrossRefGoogle ScholarPubMed
Da Costa, N, McGillivray, C and Chang, KC 2003. Postnatal myosin heavy chain isoforms in prenatal porcine skeletal muscles: insights into temporal regulation. Anatomical Record 273A 731-740.Google Scholar
Dehoux, MJM, Van Beneden, RP, Fernandez-Celemin, L, Lause, PL and Thissen, JP 2003. Induction of MafBx and Murf ubiquitin ligase mRNAs in rat skeletal muscle after LPS injection. FEBS Letters 544, 214-217.CrossRefGoogle ScholarPubMed
Delling, U, Tureckova, J, Lim, HW, De Windt, LJ, Rotwein, P and Molkentin, JD 2000. A calcineurin-NFATc3-dependent pathway regulates skeletal muscle differentiation and slow myosin heavy-chain expression. Molecular and Cellular Biology 20, 6600-6611.CrossRefGoogle ScholarPubMed
Denley, A, Cosgrove, LJ, Booker, GW, Wallace, JC and Forbes, BE 2005. Molecular interactions of the IGF system. Cytokine & Growth Factor Reviews 16, 421-439.Google Scholar
Doumit, ME, Cook, DR and Merkel, RA 1996. Testosterone up-regulates androgen receptors and decreases differentiation of porcine myogenic satellite cells in-vitro. Endocrinology 137, 1385-1394.CrossRefGoogle ScholarPubMed
Du, M, Zhu, MJ, Means, WJ, Hess, BW and Ford, SP 2005. Nutrient restriction differentially modulates the mammalian target of rapamycin signaling and the ubiquitin-proteasome system in skeletal muscle of cows and their fetuses. Journal of Animal Science 83, 117-123.CrossRefGoogle ScholarPubMed
Dunn, SE, Simard, AR, Bassel-Duby, R, Williams, RS and Michel, RN 2001. Nerve activity-dependent modulation of calcineurin signaling in adult fast and slow skeletal muscle fibers. Journal of Biological Chemistry 276, 45243-45254.Google Scholar
Dunshea, FR, D'Souza, DN, Pethick, DW, Harper, GS and Warner, RD 2005. Effects of dietary factors and other metabolic modifiers on quality and nutritional value of meat. Meat Science 71, 8-38.Google Scholar
Dwyer, CM, Stickland, NC and Fletcher, JM 1994. The influence of maternal nutrition on muscle fiber number development in the porcine fetus and on subsequent postnatal growth. Journal of Animal Science 72, 911-917.CrossRefGoogle ScholarPubMed
Erbay, E and Chen, J 2001. The mammalian target of rapamycin regulates C2C12 myogenesis via a kinase-independent mechanism. Journal of Biological Chemistry 276, 36079-36082.CrossRefGoogle ScholarPubMed
Erbay, E, Park, IH, Nuzzi, PD, Schoenherr, CJ and Chen, J 2003. IGF-II transcription in skeletal myogenesis is controlled by mTOR and nutrients. Journal of Cell Biology 163, 931-936.Google Scholar
Espinosa, L, Ingles-Esteve, J, Aguilera, C and Bigas, A 2003. Phosphorylation by glycogen synthase kinase-3β down-regulates notch activity, a link for notch and Wnt pathways. Journal of Biological Chemistry 278, 32227-32235.Google Scholar
Essén-Gustavsson, B 1993. Muscle-fiber characteristics in pigs and relationships to meat-quality parameters- review. In Pork quality: genetic and metabolic factors (ed. Puolanne, E and Demeyer, DI), pp. 140-159, CAB International, Wallingford UK.Google Scholar
Fang, SH, Nishimura, T and Takahashi, K 1999. Relationship between development and intramuscular connective tissue and toughness of pork during growth of pigs. Journal of Animal Science 77, 120-130.Google Scholar
Fraysse, B, Desaphy, JF, Pierno, S, De Luca, A, Liantonio, A, Mitolo, CI and Camerino, DC 2003. Decrease in resting calcium and calcium entry associated with slow-to-fast transition in unloaded rat soleus muscle. FASEB Journal 17, 1916-1918.Google Scholar
Fujiwara, Y, Chang, AN, Williams, AM and Orkin, SH 2004. Functional overlap of GATA-1 and GATA-2 in primitive hematopoietic development. Blood 103, 583-585.Google Scholar
Geay, Y, Bauchart, D, Hocquette, J-F and Culioli, J 2001. Effect of nutritional factors on biochemical, structural and metabolic characteristics of muscles in ruminants, consequences on dietetic value and sensorial qualities of meat. Reproduction Nutrition Development 41, 1-26.Google Scholar
Gil, F, Lopez-Albors, O, Vazquez, JM, Latorre, R, Ramirez-Zarzosa, G and Moreno, F 2001. The histochemical profiles of fibre types in porcine skeletal muscle. Histology and Histopathology 16, 439-442.Google Scholar
Giles, RH, Van Es, JH and Clevers, H 2003. Caught up in a Wnt storm: Wnt signaling in cancer. Biochimica et Biophysica Acta -Reviews on Cancer 1653, 1-24.CrossRefGoogle Scholar
Glass, DJ 2003a. Molecular mechanisms modulating muscle mass. Trends in Molecular Medicine 9, 344-350.Google Scholar
Glass, DJ 2003b. Signalling pathways that mediate skeletal muscle hypertrophy and atrophy. Nature Cell Biology 5, 87-90.Google Scholar
Glass, DJ 2005. Skeletal muscle hypertrophy and atrophy signaling pathways. International Journal of Biochemistry & Cell Biology 37, 1974-1984.Google Scholar
Goldring, K, Partridge, T and Watt, D 2002. Muscle stem cells. Journal of Pathology 197, 457-467.CrossRefGoogle ScholarPubMed
Grobet, L, Martin, LJR, Poncelet, D, Pirottin, D, Brouwers, B, Riquet, J, Schoeberlein, A, Dunner, S, Ménissier, F, Massabanda, J, Fries, R, Hanset, R and Georges, M 1997. A deletion in the bovine myostatin gene causes the double-muscled phenotype in cattle. Nature Genetics 17, 71-74.CrossRefGoogle ScholarPubMed
Haddad, F and Adams, GR 2004. Inhibition of MAP/ERK kinase prevents IGF-I induced hypertrophy in rat muscles. Journal of Applied Physiology 96, 203-210.Google Scholar
Hallauer, PL and Hastings, KEM 2002. Coregulation of fast contractile protein transgene and glycolytic enzyme expression in mouse skeletal muscle. American Journal of Physiology -Cell Physiology 282, C113-C124.CrossRefGoogle ScholarPubMed
Handschin, C, Rhee, J, Lin, J, Tarr, PT and Spiegelman, BM 2003. An autoregulatory loop controls peroxisome proliferator-activated receptor γ coactivator 1α expression in muscle. Proceedings of the National Academy of Sciences USA 100, 7111-7116.Google Scholar
Hinkle, RT, Hodge, KMB, Cody, DB, Sheldon, RJ, Kobilka, BK and Isfort, RJ 2002. Skeletal muscle hypertrophy and anti-atrophy effects of clenbuterol are mediated by the β2-adrenergic receptor. Muscle Nerve 25, 729-734.Google Scholar
Hirai, S, Matsumoto, H, Hino, N, Kawachi, H, Matsui, T and Yano, H 2007. Myostatin inhibits differentiation of bovine preadipocyte. Domestic Animal Endocrinology 32, 1-14.Google Scholar
Hocquette, JF, Ortigues-Marty, I, Pethick, D, Herpin, P and Fernandez, X 1998. Nutritional and hormonal regulation of energy metabolism in skeletal muscles of meat-producing animals. Livestock Production Science 56, 115-143.CrossRefGoogle Scholar
Holst, D, Luquet, S, Nogueira, V, Kristiansen, K, Leverve, X and Grimaldi, PA 2003. Nutritional regulation and role of peroxisome proliferator-activated receptor δ in fatty acid catabolism in skeletal muscle. Biochimica et Biophysica Acta 1633, 43-50.Google Scholar
Hornick, JL, Van Eenaeme, C, Gérard, O, Dufrasne, I and Istasse, L 2000. Mechanisms of reduced and compensatory growth. Domestic Animal Endocrinology 19, 121-132.Google Scholar
Horsley, V, Jansen, KM, Mills, ST and Pavlath, GK 2003. IL-4 acts as a myoblast recruitment factor during mammalian muscle growth. Cell 113, 483-494.Google Scholar
Horsley, V and Pavlath, GK 2002. NFAT: ubiquitous regulator of cell differentiation and adaptation. Journal of Cell Biology 156, 771-774.Google Scholar
Ji, SQ, Losinski, RL, Cornelius, SG, Frank, GR, Willis, GM, Gerrard, DE, Depreux, FS and Spurlock, ME 1998. Myostatin expression in porcine tissues: tissue specificity and developmental and postnatal regulation. American Journal of Physiology -Regulatory Integrative and Comparative Physiology 44, R1265-R1273.Google Scholar
Jiang, BH, Zheng, JZ and Vogt, PK 1998. An essential role of phosphatidylinositol 3-kinase in myogenic differentiation. Proceedings of the National Academy of Sciences USA 95, 14179-14183.CrossRefGoogle ScholarPubMed
Jope, RS 2003. Lithium and GSK-3: one inhibitor, two inhibitory actions, multiple outcomes. Trends in Pharmacological Sciences 24, 441-443.Google Scholar
Kamanga-Sollo, E, Pampusch, MS, White, ME and Dayton, WR 2003. Role of insulin-like growth factor binding protein (IGFBP)-3 in TGF-β and GDF-8 (myostatin)-induced suppression of proliferation in porcine embryonic myogenic cell cultures. Journal of Cellular Physiology 197, 225-231.Google Scholar
Kambadur, R, Sharma, M, Smith, TL and Bass, JJ 1997. Mutations in myostatin (GDF8) in double-muscled Belgian blue and Piedmontese cattle. Genome Research 7, 910-916.Google Scholar
Karlsson, AH, Klont, RE and Fernandez, X 1999. Skeletal muscle fibres as factors for pork quality. Livestock Production Science 60, 255-269.Google Scholar
Kelley, DE, Goodpaster, BH and Storlien, L 2002. Muscle triglyceride and insulin resistance. Annual Review of Nutrition 22, 325-346.Google Scholar
Kirk, SP, Oldham, JM, Jeanplong, F and Bass, JJ 2003. Insulin-like growth factor-II delays early but enhances late regeneration of skeletal muscle. Journal of Histochemistry and Cytochemistry 51, 1611-1620.Google Scholar
Klont, RE, Brocks, L and Eikelenboom, G 1998. Muscle fibre type and meat quality. Meat Science 49, S219-S229.CrossRefGoogle Scholar
Kocamis, H and Killefer, J 2002. Myostatin expression and possible functions in animal muscle growth. Domestic Animal Endocrinology 23, 447-454.Google Scholar
Koh, EH, Kim, M-S, Park, J-T, Kim, HS, Youn, J-Y, Park, H-S, Youn, JH and Lee, K-U 2003. Peroxisome proliferator-activated receptor (PPAR)-α activation prevents diabetes in OLETF rats: comparison with PPAR-γ activation. Diabetes 52, 2331-2337.CrossRefGoogle ScholarPubMed
Kutscher, EC, Luna, BC and Perry, PJ 2002. Anabolic steroids: a review for the clinician. Sports Medicine 32, 285-296.CrossRefGoogle ScholarPubMed
Langley, B, Thomas, M, Bishop, A, Sharma, M, Gilmour, S and Kambadur, R 2002. Myostatin inhibits myoblast differentiation by down-regulating MyoD expression. Journal of Biological Chemistry 277, 49831-49840.Google Scholar
LaVoie, HA 2003. The role of GATA in mammalian reproduction. Experimental Biology and Medicine 228, 1282-1290.Google Scholar
Lecker, SH, Jagoe, RT, Gilbert, A, Gomes, M, Baracos, V, Bailey, J, Price, SR, Mitch, WE and Goldberg, AL 2004. Multiple types of skeletal muscle atrophy involve a common program of changes in gene expression. FASEB Journal 18, 39-51.Google Scholar
Lee, J, Hong, F, Kwon, S, Kim, SS, Ki, DO, Kang, HS, Lee, SJ, Ha, J and Kim, SS 2002. Activation of p38 MAPK induces cell cycle arrest via inhibition of Raf/ERK pathway during muscle differentiation. Biochemical and Biophysical Research Communications 298, 765-771.CrossRefGoogle ScholarPubMed
Lefaucheur, L, Milan, D, Eolan, P and Le Callennec, C 2004. Myosin heavy chain composition of different skeletal muscles in Large White and Meishan pigs. Journal of Animal Science 82, 1931-1941.Google Scholar
Li, YP, Chen, Y, John, J, Moylan, J, Jin, B, Mann, DL and Reid, MB 2005. TNF-α acts via p38 MAPK to stimulate expression of the ubiquitin ligase atrogin1/MAFbx in skeletal muscle. FASEB Journal 19, 362-370.Google Scholar
Li, YP, Chen, Y, Li, AS and Reid, MB 2003. Hydrogen peroxide stimulates ubiquitin-conjugating activity and expression of genes for specific E2 and E3 proteins in skeletal muscle myotubes. American Journal of Physiology -Cell Physiology 285, C806-C812.Google Scholar
Lin, J, Wu, H, Tarr, PT, Zhang, CY, Wu, Z, Boss, O, Michael, LF, Puigserver, P, Isotani, E, Olson, EN, Lowell, BB, Bassel-Duby, R and Spiegelman, BM 2002. Transcriptional co-activator PGC-1α drives the formation of slow-twitch muscle fibres. Nature 418, 797-801.CrossRefGoogle Scholar
Liu, W, Thomas, SG, Asa, SL, Gonzalez-Cadavid, N, Bhasin, S and Ezzat, S 2003. Myostatin is a skeletal muscle target of growth hormone anabolic action. Journal of Clinical Endocrinology and Metabolism 88, 5490-5496.Google Scholar
Liu, Y, Cseresnyés, Z, Randall, WR and Schneider, MF 2001. Activity-dependent nuclear translocation and intranuclear distribution of NFATc in adult skeletal muscle fibers. Journal of Cell Biology 155, 27-40.Google Scholar
Lynch, GS, Cuffe, SA, Plant, DR and Gregorevic, P 2001. IGF-I treatment improves the functional properties of fast- and slow-twitch skeletal muscles from dystrophic mice. Neuromuscular Disorders 11, 260-268.Google Scholar
Maak, S, Jaesert, S, Neumann, K, Yerle, M and Von Lengerken, G 2001. Isolation of expressed sequence tags of skeletal muscle of neonatal healthy and splay leg piglets and mapping by somatic cell hybrid analysis. Animal Genetics 32, 303-307.Google Scholar
MacAulay, K, Hajduch, E, Blair, AS, Coghlan, MP, Smith, SA and Hundal, HS 2003. Use of lithium and SB-415286 to explore the role of glycogen synthase kinase-3 in the regulation of glucose transport and glycogen synthase. European Journal of Biochemistry 270, 3829-3838.Google Scholar
McCormick, RJ 1999. Extracellular modifications to muscle collagen: implications for meat quality. Poultry Science 78, 785-791.Google Scholar
McCroskery, S, Thomas, M, Maxwell, L, Sharma, M and Kambadur, R 2003. Myostatin negatively regulates satellite cell activation and self-renewal. Journal of Cell Biology 162, 1135-1147.Google Scholar
McPherron, AC and Lee, SJ 1997. Double muscling in cattle due to mutations in the myostatin gene. Proceedings of the National Academy of Sciences USA 94, 12457-12461.Google Scholar
Maccatrozzo, L, Patruno, M, Toniolo, L, Reggiani, C and Mascarello, F 2004. Myosin heavy chain 2B isoform is expressed in specialized eye muscles but not in trunk and limb muscles of cattle. European Journal of Histochemistry 48, 357-366.Google Scholar
Maltin, C, Balcerzak, D, Tilley, R and Delday, M 2003. Determinants of meat quality: tenderness. Proceedings of the Nutrition Society 62, 337-347.Google Scholar
Marchitelli, C, Savarese, MC, Crisa, A, Nardone, A, Marsan, PA and Valentini, A 2003. Double muscling in Marchigiana beef breed is caused by a stop codon in the third exon of myostatin gene. Mammalian Genome 14, 392-395.Google Scholar
Martin, B, Schneider, R, Janetzky, S, Waibler, Z, Pandur, P, Kuhl, M, Behrens, J, der Mark, K, Starzinski-Powitz, A and Wixler, V 2002. The LIM-only protein FHL2 interacts with β-catenin and promotes differentiation of mouse myoblasts. Journal of Cell Biology 159, 113-122.Google Scholar
Masuda, ES, Imamura, R, Amasaki, Y, Arai, K and Arai, N 1998. Signalling into the T-cell nucleus: NFAT regulation. Cellular Signalling 10, 599-611.Google Scholar
Mersmann, HJ 1998. Overview of the effects of β-adrenergic receptor agonists on animal growth including mechanisms of action. Journal of Animal Science 76, 160-172.Google Scholar
Mills, SE, Spurlock, ME and Smith, DJ 2003. β-Adrenergic receptor subtypes that mediate ractopamine stimulation of lipolysis. Journal of Animal Science 81, 662-668.Google Scholar
Mitin, N, Kudla, AJ, Konieczny, SF and Taparowsky, EJ 2001. Differential effects of Ras signaling through NFκB on myogenesis. Oncogene 20, 1276-1286.Google Scholar
Miura, S, Kai, Y, Ono, M and Ezaki, O 2003. Overexpression of peroxisome proliferator-activated receptor γ coactivator-1α down-regulates GLUT4 mRNA in skeletal muscles. Journal of Biological Chemistry 278, 31385-31390.Google Scholar
Morali, OG, Jouneau, A, McLaughlin, KJ, Thiery, JP and Larue, L 2000. IGFII promotes mesoderm formation. Developmental Biology 227, 133-145.Google Scholar
Morio, B, Hocquette, JF, Montaurier, C, Boirie, Y, Bouteloup-Demange, C, McCormack, C, Fellmann, N, Beaufrere, B and Ritz, P 2001. Muscle fatty acid oxidative capacity is a determinant of whole body fat oxidation in elderly people. American Journal of Physiology - Endocrinology and Metabolism 280, E143-E149.Google Scholar
Muoio, DM, Way, JM, Tanner, CJ, Winegar, DA, Kliewer, SA, Houmard, JA, Kraus, WE and Dohm, GL 2002. Peroxisome proliferator-activated receptor-α regulates fatty acid utilization in primary human skeletal muscle cells. Diabetes 51, 901-909.CrossRefGoogle ScholarPubMed
Murgia, M, Serrano, AL, Calabria, E, Pallafacchina, G, Lømo, T and Schiaffino, S 2000. Ras is involved in nerve-activity-dependent regulation of muscle genes. Nature Cell Biology 2, 142-147.Google Scholar
Musarò, A, McCullagh, JA, Naya, FJ, Olson, EN and Rosenthal, N 1999. IGF-1 induces skeletal myocyte hypertrophy through calcineurin in association with GATA-2 and NF-ATc1. Nature 400, 581-585.Google Scholar
Musarò, A, McCullagh, K, Paul, A, Houghton, L, Dobrowolny, G, Molinaro, M, Barton, ER, Sweeney, HL and Rosenthal, N 2001. Localized Igf-1 transgene expression sustains hypertrophy and regeneration in senescent skeletal muscle. Nature Genetics 27, 195-200.Google Scholar
Naya, FJ, Mercer, B, Shelton, J, Richardson, JA, Williams, RS and Olson, EN 2000. Stimulation of slow skeletal muscle fiber gene expression by calcineurin in vivo. Journal of Biological Chemistry 275, 4545-4548.Google Scholar
Naya, FJ and Olson, E 1999. MEF2: a transcriptional target for signaling pathways controlling skeletal muscle growth and differentiation. Current Opinion in Cell Biology 11, 683-688.Google Scholar
Nissen, PM, Danielsen, VO, Jorgensen, PF and Oksbjerg, N 2003. Increased maternal nutrition of sows has no beneficial effects on muscle fiber number or postnatal growth and has no impact on the meat quality of the offspring. Journal of Animal Science 81, 3018-3027.Google Scholar
Novak, A and Dedhar, S 1999. Signaling through β-catenin and Lef/Tcf. Cellular and Molecular Life Sciences 56, 523-537.Google Scholar
Oberkofler, H, Esterbauer, H, Linnemayr, V, Strosberg, AD, Krempler, F and Patsch, W 2002. Peroxisome proliferator-activated receptor (PPAR) γ coactivator-1 recruitment regulates PPAR subtype specificity. Journal of Biological Chemistry 277, 16750-16757.Google Scholar
Oka, A, Iwaki, F, Dohgo, T, Ohtagaki, S, Noda, M, Shiozaki, T, Endoh, O and Ozaki, M 2002. Genetic effects on fatty acid composition of carcass fat of Japanese Black Wagyu steers. Journal of Animal Science 80, 1005-1011.CrossRefGoogle ScholarPubMed
Ooi, PT, Da Costa, N, Edgar, J and Chang, KC 2006. Porcine congenital splayleg is characterised by muscle fibre atrophy associated with relative rise in MAFbx and fall in P311 expression. BMC Veterinary Research 2, 23.Google Scholar
Pallafacchina, G, Calabria, E, Serrano, AL, Kalhovde, JM and Schiaffino, S 2002. A protein kinase B-dependent and rapamycin-sensitive pathway controls skeletal muscle growth but not fiber type specification. Proceedings of the National Academy of Sciences USA 99, 9213-9218.CrossRefGoogle Scholar
Parkington, JD, Siebert, AP, LeBrasseur, NK and Fielding, RA 2003. Differential activation of mTOR signaling by contractile activity in skeletal muscle. American Journal of Physiology -Regulatory Integrative and Comparative Physiology 285, R1086-R1090.Google Scholar
Paul, AC and Rosenthal, N 2002. Different modes of hypertrophy in skeletal muscle fibers. Journal of Cell Biology 156, 751-760.Google Scholar
Petropoulos, H and Skerjanc, IS 2002. β-Catenin is essential and sufficient for skeletal myogenesis in P19 Cells. Journal of Biological Chemistry 277, 15393-15399.CrossRefGoogle ScholarPubMed
Pette, D and Staron, RS 2000. Myosins isoforms, muscle fiber types, and transitions. Microscopy Research and Technique 50, 500-509.Google Scholar
Philip, B, Lu, Z and Gao, Y 2005. Regulation of GDF-8 signaling by the p38 MAPK. Cellular Signalling 17, 365-375.Google Scholar
Picard, B, Gagniere, H, Geay, Y, Hocquette, J-F and Robelin, J 1995. Study of the influence of age and weaning on the contractile and metabolic characteristics of bovine muscle. Reproduction Nutrition Development 35, 71-84.Google Scholar
Picard, B, Lefaucheur, L, Berri, C and Duclos, MJ 2002. Muscle fibre ontogenesis in farm animal species. Reproduction Nutrition Development 42, 415-431.Google Scholar
Polesskaya, A, Seale, P and Rudnicki, MA 2003. Wnt signaling induces the myogenic specification of resident CD45+ adult stem cells during muscle regeneration. Cell 113, 841-852.Google Scholar
Prunier, C, Pessah, M, Ferrand, N, Seo, SR, Howe, P and Atfi, A 2003. The oncoprotein Ski acts as an antagonist of transforming growth factor-β signaling by suppressing Smad2 phosphorylation. Journal of Biological Chemistry 278, 26249-26257.Google Scholar
Rao, A, Luo, C and Hogan, PG 1997. Transcription factors of the NFAT family:regulation and function. Annual Review of Immunology 15, 707-747.Google Scholar
Rehfeldt, C, Kuhn, G, Vanselow, J, Fürbass, R, Fiedler, I, Nürnberg, G, Clelland, AK, Stickland, NC and Ender, K 2001. Maternal treatment with somatotropin during early gestation affects basic events of myogenesis in pigs. Cell and Tissue Research 306, 429-440.Google Scholar
Relaix, F 2006. Skeletal muscle progenitor cells: from embryo to adult. Cellular and Molecular Life Sciences 63, 1221-1225.Google Scholar
Rommel, C, Bodine, SC, Clarke, BA, Rossman, R, Nunez, L, Stitt, TN, Yancopoulos, GD and Glass, DJ 2001. Mediation of IGF-1-induced skeletal myotube hypertrophy by PI(3)K/Akt/mTOR and PI(3)K/Akt/GSK3 pathways. Nature Cell Biology 3, 1009-1013.Google Scholar
Rommel, C, Clarke, BA, Zimmermann, S, Nuñez, L, Rossman, R, Reid, K, Moelling, K, Yancopoulos, GD and Glass, DJ 1999. Differentiation stage-specific inhibition of the Raf-MEK-ERK pathway by Akt. Science 286, 1738-1744.Google Scholar
Rothermel, BA, McKinsey, TA, Vega, RB, Nicol, RL, Mammen, P, Yang, J, Antos, CL, Shelton, JM, Bassel-Duby, R, Olson, EN and Williams, RS 2001. Myocyte-enriched calcineurin-interacting protein, MCIP1, inhibits cardiac hypertrophy in vivo. Proceedings of the National Academy of Sciences USA 98, 3328-3333.CrossRefGoogle ScholarPubMed
Ryall, JG, Gregorevic, P, Plant, DR, Sillence, MN and Lynch, GS 2002. β2-Agonist fenoterol has greater effects on contractile function of rat skeletal muscles than clenbuterol. American Journal of Physiology -Regulatory Integrative and Comparative Physiology 283, R1386-R1394.Google Scholar
Sakamoto, K, Aschenbach, WG, Hirshman, MF and Goodyear, LJ 2003. Akt signaling in skeletal muscle: regulation by exercise and passive stretch. American Journal of Physiology - Endocrinology and Metabolism 285, E1081-E1088.Google Scholar
Sakuma, K, Nishikawa, J, Nakao, R, Watanabe, K, Totsuka, T, Nakano, H, Sano, M and Yasuhara, M 2003. Calcineurin is a potent regulator for skeletal muscle regeneration by association with NFATc1 and GATA-2. Acta Neuropathologica 105, 271-280.Google Scholar
Sandri, M, Sandri, C, Gilbert, A, Skurk, C, Calabria, E, Picard, A, Walsh, K, Schiaffino, S, Lecker, SH and Goldberg, AL 2004. Foxo transcription factors induce the atrophy-related ubiquitin ligase atrogin-1 and cause skeletal atrophy. Cell 117, 399-412.Google Scholar
Sant'ana Pereira, JA, Wessels, A, Nijtmans, L, Moorman, AF and Sargeant, AJ 1995. New method for the accurate characterization of single human skeletal muscle fibres demonstrates a relation between mATPase and MyHC expression in pure and hybrid fibre types. Journal of Muscle Research and Cell Motility 16, 21-34.CrossRefGoogle ScholarPubMed
Schiaffino, S and Reggiani, C 1996. Molecular diversity of myofibrillar proteins: gene regulation and functional significance. Physiological Reviews 76, 371-423.Google Scholar
Schiaffino, S and Reggiani, C 1994. Myosin isoforms in mammalian skeletal muscle. Journal of Applied Physiology 77, 493-501.Google Scholar
Schulz, RA and Yutzey, KE 2004. Calcineurin signaling and NFAT activation in cardiovascular and skeletal muscle development. Developmental Biology 266, 1-16.Google Scholar
Semsarian, C, Wu, MJ, Ju, YK, Marcinec, T, Yeoh, T, Allen, DG, Harvey, RP and Graham, RM 1999. Skeletal muscle hypertrophy is mediated by a Ca2+-dependent calcineurin signalling pathway. Nature 400, 576-581.Google Scholar
Shi, DL, Bourdelas, A, Umbhauer, M and Boucaut, JC 2002. Zygotic Wnt/β-catenin signaling preferentially regulates the expression of Myf5 gene in the mesoderm of xenopus. Developmental Biology 245, 124-135.Google Scholar
Sillence, MN, Munn, KJ and Campbell, RG 2002. Manipulation of growth in pigs through treatment of the neonate with clenbuterol and somatotropin. Journal of Animal Science 80, 1852-1862.Google Scholar
Sinha-Hikim, I, Roth, SM, Lee, MI and Bhasin, S 2003. Testosterone-induced muscle hypertrophy is associated with an increase in satellite cell number in healthy, young men. American Journal of Physiology - Endocrinology and Metabolism 285, E197-E205.Google Scholar
Soeta, C, Suzuki, M, Suzuki, S, Naito, K, Tachi, C and Tojo, H 2001. Possible role for the c-ski gene in the proliferation of myogenic cells in regenerating skeletal muscles of rats. Development Growth and Differentiation 43, 155-164.CrossRefGoogle ScholarPubMed
Spangenburg, EE and Booth, FW 2003. Molecular regulation of individual skeletal muscle fibre types. Acta Physiologica Scandinavica 178, 413-424.CrossRefGoogle ScholarPubMed
Steelman, CA, Recknor, JC, Nettleton, D and Reecy, JM 2006. Transcriptional profiling of myostatin-knockout mice implicates Wnt signaling in postnatal skeletal muscle growth and hypertrophy. FASEB Journal 20, 580-582.Google Scholar
Stitt, TN, Drujan, D, Clarke, BA, Panaro, F, Timofeyva, Y, Kline, WO, Gonzalez, M, Yancopoulos, GD and Glass, DJ 2004. The IGF-1/PI3K/Akt pathway prevents expression of muscle atrophy-induced ubiquitin ligases by inhibiting FOXO transcription factors. Molecular Cell 14, 395-403.Google Scholar
Sugiura, T, Sio, SO, Shuntoh, H and Kuno, T 2001. Molecular genetic analysis of the calcineurin signaling pathways. Cellular and Molecular Life Sciences 58, 278-288.Google Scholar
Sutrave, P, Kelly, AM and Hughes, SH 1990. Ski can cause selective growth of skeletal muscle in transgenic mice. Genes and Development 4, 1462-1472.Google Scholar
Sutrave, P, Leferovich, JM, Kelly, AM and Hughes, SH 2000. The induction of skeletal muscle hypertrophy by a ski transgene is promoter-dependent. Gene 241, 107-116.Google Scholar
Swoap, SJ, Hunter, RB, Stevenson, EJ, Felton, HM, Kansagra, NV, Lang, JM, Esser, KA and Kandarian, SC 2000. The calcineurin-NFAT pathway and muscle fiber-type gene expression. American Journal of Physiology -Cell Physiology 279, C915-C924.Google Scholar
Tajbakhsh, S and Buckingham, M 2000. The birth of muscle progenitor cells in the mouse: spatiotemporal considerations. Current Topics in Developmental Biology 228, 225-268.Google Scholar
Tesseraud, S, Metayer, S, Duchene, S, Bigot, K, Grizard, J and Dupont, J 2007. Regulation of protein metabolism by insulin: value of different approaches and animal models. Domestic Animal Endocrinology (in press).Google Scholar
Thomas, M, Langley, B, Berry, C, Sharma, M, Kirk, S, Bass, J and Kambadur, R 2000. Myostatin, a negative regulator of muscle growth, functions by inhibiting myoblast proliferation. Journal of Biological Chemistry 275, 40235-40243.Google Scholar
Ueki, N and Hayman, MJ 2003. Direct interaction of Ski with either Smad3 or Smad4 is necessary and sufficient for Ski-mediated repression of transforming growth factor-β signaling. Journal of Biological Chemistry 278, 32489-32492.Google Scholar
Van Rooij, E, Doevendans, PA, CrijnsHJGM, HJGM,, Heeneman, S, Lips, DJ, Van Bilsen, M, Williams, RS, Olson, EN, Bassel-Duby, R, Rothermel, BA, De Windt, LJ, et al. , 2004. MCIP1 overexpression suppresses left ventricular remodeling and sustains cardiac function after myocardial infarction. Circulation Research 94, 18e-26.Google Scholar
Vollenweider, P 2003. Insulin resistant states and insulin signaling. Clinical Chemistry and Laboratory Medicine 41, 1107-1119.Google Scholar
Vyas, DR, Spangenburg, EE, Abraha, TW, Childs, TE and Booth, FW 2002. GSK-3β negatively regulates skeletal myotube hypertrophy. American Journal of Physiology -Cell Physiology 283, C545-C551.Google Scholar
Wang, Y-X, Lee, C-H, Tiep, S, Yu, RT, Ham, J, Kang, H and Evans, RM 2003. Peroxisome-proliferator-activated receptor δ activates fat metabolism to prevent obesity. Cell 113, 159-170.Google Scholar
Weiss, A and Leinwand, LA 1996. The mammalian myosin heavy chain gene family. Annual Review of Cell and Developmental Biology 12, 417-439.Google Scholar
Wigmore, PMC and Stickland, NC 1983. Muscle development in large and small pig fetuses. Journal of Anatomy 2, 235-245.Google Scholar
Wilson, EM, Hsieh, MM and Rotwein, P 2003. Autocrine growth factor signaling by insulin-like growth factor-II mediates MyoD-stimulated myocyte maturation. Journal of Biological Chemistry 278, 41109-41113.Google Scholar
Wolfe, RR 2006. The underappreciated role of muscle in health and disease. American Journal of Clinical Nutrition 84, 475-482.Google Scholar
Wood, JD, Enser, M, Fisher, AV, Nute, GR, Richardson, RI and Sheard, PR 1999. Manipulating meat quality and composition. Proceedings of the Nutrition Society 58, 363-370.Google Scholar
Wood, JD, Nute, GR, Richardson, RI, Whittington, FM, Southwood, O, Plastow, G, Mansbridge, R, da Costa, N and Chang, KC 2004. Effects of breed, diet and muscle on fat deposition and eating quality in pigs. Meat Science 67, 651-667.Google Scholar
Wood, JD, Richardson, RI, Nute, GR, Fisher, AV, Campo, MM, Kasapidou, E, Sheard, PR and Enser, M 2003. Effects of fatty acids on meat quality: a review. Meat Science 66, 21-32.Google Scholar
Wu, G, Bazer, FW, Wallace, JM and Spencer, TE 2006. Intrauterine growth retardation: implications for the animal sciences. Journal of Animal Science 84, 2316-2337.Google Scholar
Wu, H, Naya, FJ, McKinsey, TA, Mercer, B, Shelton, JM, Chin, ER, Simard, AR, Michel, RN, Bassel-Duby, R, Olson, EN and Williams, RS 2000a. MEF2 responds to multiple calcium-regulated signals in the control of skeletal muscle fiber type. EMBO Journal 19, 1963-1973.Google Scholar
Wu, H, Rothermel, B, Kanatous, S, Rosenberg, P, Naya, FJ, Shelton, M, Hutcheson, KA, DiMaio, M, Olson, EN, Bassel-Duby, R and Williams, RS 2001. Activation of MEF2 by muscle activity is mediated through a calcineurin-dependent pathway. EMBO Journal 20, 6414-6423.Google Scholar
Wu, YZ, Crumley, RL and Caiozzo, VJ 2000b. Are hybrid fibers a common motif of canine laryngeal muscles?: single-fiber analyses of myosin heavy-chain isoform composition. Archives of Otolaryngology - Head and Neck Surgery 126, 865-873.Google Scholar
Wu, Z, Woodring, PJ, Bhakta, KS, Tamura, K, Wen, F, Feramisco, JR, Karin, M, Wang, JYJ and Puri, PP 2000c. p38 and extracellular signal-regulated kinases regulate the myogenic program at multiple steps. Molecular and Cellular Biology 20, 3951-3964.Google Scholar
Yang, J, Rothermel, B, Vega, RB, Frey, N, McKinsey, TA, Olson, EN, Bassel-Duby, R and Williams, RS 2000. Independent signals control expression of the calcineurin inhibitory proteins MCIP1 and MCIP2 in striated muscles. Circulation Research 87, 61-68.Google Scholar
Yang, SY and Goldspink, G 2002. Different roles of the IGF-I Ec peptide (MGF) and mature IGF-I in myoblast proliferation and differentiation. FEBS Letters 522, 156-160.Google Scholar
Yimlamai, T, Dodd, SL, Borst, SE and Park, S 2005. Clenbuterol induces muscle-specific attenuation of atrophy through effects on the ubiquitin-proteasome pathway. Journal of Applied Physiology 99, 71-80.Google Scholar
Yu, YH, Liu, BH, Mersmann, HJ and Ding, ST 2006. Porcine peroxisome proliferator-activated receptor γ induces transdifferentiation of myocytes into adipocytes. Journal of Animal Science 84, 2655-2665.Google Scholar
Zierath, JR and Hawley, JA 2006. Skeletal muscle fiber type: influence on contractile and metabolic properties. PLoS Biology 2, 1523-1527.Google Scholar
Zimmermann, S and Moelling, K 1999. Phosphorylation and regulation of Raf by Akt (protein kinase B). Science 286, 1741-1744.Google Scholar
Zorzano, A, Kaliman, P, Guma, A and Palacin, M 2003. Intracellular signals involved in the effects of insulin-like growth factors and neuregulins on myofibre formation. Cellular Signalling 15, 141-149.Google Scholar
Figure 0

Figure 1 Categories of skeletal muscle fibre phenotype.

Figure 1

Figure 2 The IGF-induced Erk-MAPK proliferation signalling pathway of hyperplasia. IGF-1, IGF-2 or insulin binding to tyrosine kinase receptor triggers a phosphorylation cascade that leads to transcriptional and protein modifications, culminating in increase cell proliferation. Integrin can also activate the Erk-MAPK pathway independent of tyrosine kinase receptor binding. Note that activated Ras can also activate the PI3K pathway.

Figure 2

Figure 3 The IGF-induced PI3K-Akt1 signalling pathway of muscle differentiation and hypertrophy. It affects a range of cellular processes, including the promotion of protein synthesis, glycogen synthesis, and cell proliferation and differentiation, that culminate in muscle hypertrophy. mTOR can be activated by amino acids (nutrients) and induce local IGF-2 production which, in turn, acts in a positive feedback manner. The PI3K cascade can be independently activated by exercise and by integrin stimulation (see also integrin stimulation of Erk-MAPK pathway, Figure 2). Activated Akt1 via inhibiting Raf and MEK phosphorylation is involved in inhibitory crosstalk with the Erk-MAPK pathway. Selected abbreviations: PI3K =  phosphatidylinositol 3′-kinase, PDK1 =  phosphatidylinositol-dependent protein kinase 1, GSK3β =  glycogen synthase kinase 3 β, mTOR =  mammalian target of rapamycin, GS =  glycogen synthase, and p70S6K =  ribosomal protein S6 kinase. Note that Akt, GSK3β, mTOR, p70S6k and PHAS-1 are serine/threonine kinases.

Figure 3

Figure 4 The p38 MAPK stress response pathway of cell arrest and atrophy. It is activated by a variety of stimuli, including exercise and activated Ras, and through different receptor families. It mediates a range of effects: cell arrest, differentiation, atrophy as well as mitochondrial biogenesis. Additionally, TGFβ receptors readily signal through Smad proteins which also lead to cell arrest. Ski functions to inhibit Smad signalling thus relieving cell arrest and promoting proliferation. Selected abbreviations: IL-1 =  interleukin 1, TLRs =  toll-like receptors, TNF α =  tumour necrosis factor α, and TAK =  TGFβ-activated kinase.

Figure 4

Figure 5 Typing of porcine muscle fibres. Serial sections of the psoas of a 22-week-old pig to illustrate the combined use of histochemistry, immunocytochemistry and in situ hybridisation to identify the four major post-natal MyHC fibre types. Succinate dehydrogenase (SDH) staining (a), sudan black staining (b), NOQ7.5.4D (slow) monoclonal antibody binding (c), and in situ hybridisations with MyHC isoform-specific fast 2a probe (d), fast 2b (e) and fast 2x probe (f) were performed. SDH and Sudan Black positive fibres closely mirrored each other and corresponded mainly to slow and fast 2a fibres. Note the presence of inter-fibre adipose tissue on panel b.

Figure 5

Table 1 Metabolic profiles of MyHC fibre types

Figure 6

Figure 6 Summary actions of signalling factors and pathways on fibre size, fibre number and fibre types. p38 MAPK activation can result in a range of phenotypic outcome depending on the nature of activation. The proposed inhibition of myostatin by Ski has not been formally demonstrated.