Muscle Cell Growth and Development (2024)

RONALD E. ALLEN

Skeletal muscle from domestic animals is a major source of high-quality protein in the human diet. Past technological advances in production of animal muscle protein have been based on empirical and fundamental biological research. Future technological advances, however, are less likely to occur unless research is firmly grounded in the basic biology. of muscle and animal growth. The primary function of this paper is to review information about the structure and composition of muscle, muscle differentiation and development, and key elements of protein metabolism as they relate to muscle growth. It also describes current areas of active research interest and speculates on applications of new research knowledge and future research needs.

Muscle Cell Structure and Composition

The differentiated muscle cell in postnatal muscle is the muscle fiber, a highly specialized, long, cylindrical cell that can range in diameter from 10 to 100 mm and in length from millimeters up to many centimeters. The primary differences in fibers of different species are fiber length and number of fibers per muscle. Each fiber is surrounded by a 7.5- to 10-nm-thick plasmalemma, called the sarcolemma. The sarcolemma is a lipid bilayer like the cell membranes of other cells and has a lipid composition of roughly 60 percent protein, 20 percent phospholipid, and 20 percent cholesterol. Surrounding the sarcolemma is the basal lamina, or basem*nt membrane. This somewhat amorphous structure, 50 to 70 nm thick, is composed of mucopolysaccharides and collagen (types III and V). The cell membrane of muscle has a specialized structure—the motor endplate—which accommodates interaction with an axon from a motoneuron. In addition, the membrane maintains an electrical potential that is propagated from the motor endplate, down the membrane, and finally into the cell by a complex set of invagin*tions that form the transverse tubular system.

Muscle fibers contain the major organelles present in most cells. The most striking difference between muscle cells and the majority of other cells is their multinucleated nature. Depending on its size, an individual fiber may contain hundreds of nuclei. They are found just beneath the sarcolemma and seem to be randomly distributed along the length of the fiber. Mitochondria are present between the contractile elements of muscle; their concentration varies with the metabolic activity of the particular fiber. Ribosomes are dispersed within the cytoplasm, but very few are associated with endoplasmic reticulum, primarily because muscle fibers synthesize few secreted proteins. The endoplasmic reticulum in muscle has formed a specialized set of membrane structures called the sarcoplasmic reticulum. The primary function of this structure is regulation of free calcium ion concentration. When free calcium ion concentration is maintained below approximately 0.1 mM, contraction does not occur. But when the membrane is depolarized, the action potential reaches the interior of the cell through the transverse tubular system, calcium is released from the sarcoplasmic reticulum, the concentration approaches 1 mM, and contraction is activated. Lysosomes are not readily seen in muscle fibers, although lysosomal enzymes are present. The lysosomes are most likely sequestered in the sarcoplasmic reticulum.

By far the most unique subcellular aspect of muscle fibers is the contractile machinery, the myofibril. This is an aggregation of 12 to 14 proteins into highly organized contractile threads that are insoluble at the ionic strength of the cytoplasm in muscle cells. It is noteworthy that this specialized set of proteins constitutes about 55 percent of the total protein in muscle. Consequently, many developmental studies of muscle have focused on myofibrillar protein gene expression and synthesis, which are discussed later in this paper.

Myofibrils are composed of two main classes of filaments: thick filaments and thin filaments. Thick filaments measure approximately 15 nm by 1,500 nm. The major protein in thick filaments is myosin, which has the active site that hydrolyzes adenosine triphosphate (ATP) and the site that binds to actin in the thin filament. The thin filament is roughly 6 nm by 1,000 nm and is composed of actin, which forms the beaded backbone of the filament, and tropomyosin and troponin, which perform regulatory functions. At one end, thin filaments insert into a protein lattice called the Z-line; at the other end, they overlay with thick filaments in a hexagonal array. Additional small-diameter filament systems are present within myofibrils to provide an elastic component. Also, an intermediate-diameter filament system, found outside the periphery of the myofibril, links adjacent myofibrils and maintains their contractile units in register. Specific details of the ultrastructure of myofibrils and the biochemical properties of this interdigitating array of filaments can be found in Goll et al. (1984).

These features of muscle cells are common to all skeletal muscle fibers, but specific fibers have differentiated somewhat depending on their purpose. Some populations of fibers are primarily responsible for rapid contractions on an intermittent basis, while others have slower contraction speed and sustain contractile activity over extended periods of time. Muscle fiber types have been described extensively in many species; and their biochemical, physiological, and morphological differences are significant to problems of muscle growth and meat quality. A generalized scheme for describing fiber types classifies them on the basis of their contraction speed and on the energy metabolism pathways primarily used to provide energy for contraction. Peter et al. (1972) provided one of the most descriptive classification systems by grouping fibers into three general categories. Fibers that were dependent on oxidative metabolism and had slower contraction speeds were classified as slow-twitch, oxidative fibers (so). Fibers with faster contraction times that were dependent on anaerobic, or glycolytic, energy metabolism pathways were termed fast-twitch, glycolytic fibers (FG). A third broad category contained fast-twitch fibers that had glycolytic metabolic capabilities but also a significant capacity for oxidative metabolism; these were termed fast-twitch, oxidative-glycolytic fibers (FOG).

Contraction speed is correlated with myosin adenosine triphosphatase (ATPase) activity and, therefore, with the particular myosin isozymes synthesized by the fiber. Other myofibrillar protein isoform variations may also be associated with contractile properties. The complexity and degree of development of the sarcoplasmic reticulum, t-tubule system, and neuromuscular junctions have all been associated with contraction speed and fiber class. As expected, mitochondrial content and glycolytic enzyme content vary, among fiber types, as do energy substrates such as glycogen and triglyceride. Aspects of fiber type variation that affect muscle growth include the notable differences in fiber size that generally correlate with muscle fiber type. SO fibers are smaller in diameter than FG fibers, and FOG fibers tend to be intermediate in size. Smaller fiber diameters may facilitate efficient gas exchange in oxidative fibers. In addition, SO fibers tend to have higher nuclei concentrations and, therefore, lower protein concentrations per nucleus. Satellite cell frequency, however, is reportedly higher for SO fibers (Kelly, 1978b). Because individual muscles vary in fiber type composition, factors that differentially affect the development or growth of specific fiber types can result in alterations in muscle mass (for example, the transition from FG to FOG that can accompany aerobic conditioning). Reductions in fiber diameter and, consequently, muscle mass would be expected. Alterations in gene expression and in quantitative aspects of protein metabolism that are responsible for such fiber type transitions are poorly understood.

Chemical composition of muscle tissue can be quite variable, and the primary source of variation is intramuscular adipose tissue. It is clear that most of the variation in major constituents is minimized when expressed on a fat-free basis. Some compositional variation can be found in association with aging, but, in general, it is attributable to changes in moisture content. Skeletal muscle from very young animals has a high moisture content that decreases with maturity. As a result, protein concentration increases with maturity. Subtle changes in other constituents, such as glycogen, can vary among muscles and species, but these differences may not have major nutritional significance when considering the composition of muscle as a food.

The primary lipid fraction contributing to muscle tissue variation is triglyceride, which is stored in adipocytes within the muscle. These depositions are commonly referred to as marbling, and within the range of marbling found in the longissimus muscle of beef, the ether-extractable lipid (primarily triglyceride) varies from 1.77 to 10.42 percent on a wet weight basis (Savell et al., 1986).

Cholesterol content, on the other hand, is less variable. This can best be understood in light of its role in muscle tissue. Cholesterol is an integral part of cell membranes, mainly the plasma membrane. On a tissue basis across maturity groups and marbling contents within maturity groups, cholesterol content of beef muscle does not vary (Stromer et al., 1966). In addition, the amount of cholesterol per gram of whole steak was not significantly different among the five yield grades examined by Rhee et al. (1982). Furthermore, neither breed type nor nutritional background affected cholesterol content of lean muscle tissue in beef cows (Eichhorn et al., 1986). It is possible to find variation in cholesterol content of meat, however, because adipose tissue tends to have a higher cholesterol concentration than do muscle fibers. Consequently, variations in the amount of subcutaneous or inter-muscular fat consumed with the lean portion can alter cholesterol intake. It has been calculated that 37 to 56 percent of the cholesterol in a cooked rib steak of beef originates from subcutaneous and inter-muscular adipose tissue (Rhee et al., 1982).

In looking only at muscle cells, however, significant variations in cholesterol content have not been seen, even among most of the species used for muscle foods (Reiser, 1975; Watt and Merrill, 1963). This is also true for the amino acid composition of muscle. The majority of muscle cell proteins are myofibrillar and are very highly conserved across species. In addressing topics such as alteration of tissue composition to enhance nutritional quality, it is important to keep in mind that the biology of the animal or tissue must come first. Our ability to manipulate cells in animals has both physiological limits and ramifications.

Muscle Fiber Development

Prenatal Development

Myogenesis originates in cells of the embryonic mesoderm and apparently follows a similar course in all species examined. Perhaps the most detailed descriptions come from studies of human (Hauschka, 1974) and chick (White et al., 1975) embryo development. In the human, no apparent organization is noted in the limb mesoderm on day 28 of development, but by day 43 loose connective tissue cell regions and compact myogenic cell regions are visible. By day 45 the first small multinucleated myotubes (the precursors of muscle fibers) have formed; by day 50 the general organization of major muscles and bones is essentially complete. Beyond this point, the rate of muscle histogenesis occurs at different rates between and within individual muscles. In the gastrocnemius on day 62, well-developed, my-ofibril-containing muscle fibers are present, but the majority of cells are still mononucleated. This population decreases to about 50 percent of the total by day 72, while fibers increase two- to threefold. During the next 2 weeks, fiber formation proceeds rapidly, with the percentage of mononucleated cells diminishing to 20 percent by day 95 and further decreasing to the point that only a few single cells persist in association with fibers by day 146.

In other vertebrate species, comparable developmental patterns are discernible. One striking observation in rat and chick muscle is the development of two populations of fibers (Kelly and Zacks, 1969; McLennan, 1983). The ''primary fibers'' develop early and are surrounded by closely associated mononucleated cells. In the chick embryo, "secondary fiber" formation proceeds rapidly after about 12 days of development until most of the mononucleated cell population is exhausted and fiber formation is complete. This occurs before hatching in the chick and before birth in most mammals. A similar biphasic developmental pattern has been documented in fetal lamb skeletal muscle (Ashmore et al., 1972). In general, fiber formation is complete near the time of birth.

The study of myogenesis focuses on the muscle development process and has centered around efforts to unravel myogenic lineages and the mechanisms responsible for alterations in the synthetic programs of muscle cells that lead to the formation of fibers and the expression of muscle-specific cell characteristics. One of the most important initial observations on the mechanisms of myogenesis came from a series of experiments reported by Stockdale and Holtzer (1961) that directly demonstrate that multinucleated myotubes arise from the fusion of mononucleated myogenic cells (myoblasts). Furthermore, only mononucleated cells have the ability to proliferate; the nuclei in myotubes cannot replicate their DNA and divide. Consequently, the transition from a proliferating myoblast to a nonproliferating myotube that can synthesize muscle-specific macromolecules represents the terminal step in muscle differentiation.

There now appear to be several different types of myogenic cells that are actively proliferating and differentiating during specific periods of development. Their collective developmental patterns are responsible for the general pattern of muscle histogenesis. At least two broad types and four subtypes of myogenic cells have been identified by White et al. (1975), based on the in vitro morphology and medium requirements of cloned myogenic cells from various stages of embryo development. One general type is the early muscle-colony-forming cell, which predominates in early development; the colonies are noted for having small, thick myotubes with few nuclei. In contrast, the predominant form of myogenic cells in later periods of development form colonies in vitro that are extensively fused and contain large myotubes with many nuclei; these are the late muscle-colony-forming cells. Miller and Stockdale (1986) have identified four types of myogenic cells based on the presence of specific isoforms of the myosin heavy chains present in early and late muscle-colony-forming cells.

Early and late classes of cells appear to be distinct, since they can maintain their class-specific characteristic when subcloned up to five times, until proliferative senescence (Rutz and Hauschka, 1982). Additional experiments reported by Seed and Hauschka (1984) have shown that transplanting limb buds at various stages results in the absence of late myogenic cells in the transplant, even though the early class of muscle-colony-forming cells was present. The late class apparently migrates into the limb bud from the somite at a later stage than the early class and, furthermore, does not appear to descend from the early class, in agreement with the previous in vitro experiments (Rutz and Hauschka, 1982). The appearance of early and late muscle-colony-forming cells appears to correlate well with the anatomical appearance of primary and secondary fibers that are formed during development. Different myogenic classes of cells are further implicated in the formation of primary and secondary fibers because the in vivo formation of secondary fibers is nerve-dependent (McLennan, 1983), as is the in vitro development of fibers from one of the later muscle-colony-forming types (Bonner and Adams, 1982).

An additional class of myogenic cells, or branch of the myogenic lineage, is the satellite cell, which is discussed further in the subsection on postnatal development.

As mentioned previously, a striking transition takes place in muscle development with the differentiation of mononucleated myoblasts into multinucleated myofibers. This terminal step in differentiation is accompanied by the cessation of proliferation and the expression of genes responsible for the muscle phenotype. For many years, there were two general theories to explain myogenesis. The first postulated that a major reorganization in gene expression took place in specific mitotic cycles, and the resultant daughter cells had protein synthesis capabilities that differed from those of the mother cell. This special cell cycle was referred to as a "quantal" cell cycle (Holtzer and Bischoff, 1970). This theory has now been expanded to hypothesize that a fixed number of cell divisions occur between the stem cell compartment to the terminally differentiated, fusion-competent myoblast compartment (Quinn et al., 1984). Key to this description of myogenesis is the "commitment" step of myoblasts to withdraw from the cell cycle, fuse, and initiate the synthesis of muscle-specific proteins.

In contrast, a second theory, of myogenesis (Buckley and Konigsberg, 1974) was based on a model that predicted that myoblasts remaining in the G1 phase of the cell cycle had an increasing probability of fusion that resulted in permanent withdrawal from the cell cycle and the initiation of muscle protein synthesis. The probability of remaining in the cell cycle or fusing depended on the presence or absence of environmental factors that stimulate these activities. In this model, withdrawal from the cycle and initiation of muscle gene expression was thought to be the result of the fusion process itself.

A current, and more likely, explanation encompasses elements of both the original theories. It appears that during the early part of the G1 phase of the cell cycle, proliferating myoblasts have the option of continuing to proliferate or of differentiating and fusing into myotubes (Nadal-Ginard, 1978). The commitment to withdraw from the cell cycle is made before fusion, not as a result of fusion. This commitment, however, depends on the presence of growth-stimulating factors in the environment (probably mitogens) that keep myoblasts in the cell cycle. For many years, it appeared that withdrawal from the cell cycle, fusion, and expression of the muscle phenotype were coupled events; recent experiments with a temperature-sensitive mutant of the muscle cell line L6E9 have cast doubts on the obligatory relationship of these events. In experiments with wild-type and mutant L6E9 myoblasts, Nguyen et al. (1983) demonstrated that muscle-specific isoforms of certain myofibrillar proteins could be induced in the mutant cells under conditions that did not permit commitment to withdrawal from the cell cycle. In fact, these cells could be stimulated to reenter the cell cycle even after induction of myofibrillar protein synthesis. Additional experimentation with wild-type L6E9 myoblasts arrested in a low-calcium medium indicated that induction of myofibrillar protein synthesis occurred in cells that could subsequently be stimulated to synthesize DNA and divide. Reentry into the cell cycle, however, resulted in a rapid cessation of myofibrillar protein synthesis and degradation of existing muscle-specific messenger RNAs (Nadal-Ginard et al., 1984). Similar experiments were reported with primary cultures of quail embryo muscle that were arrested in a low-calcium medium (Devlin and Konigsberg, 1983). Apparently, induction of the gene expression transitions leading to the muscle phenotype can be uncoupled from permanent withdrawal from the cell cycle. In normal muscle development, however, the commitment to withdraw from the cell cycle and the induction of the muscle phenotype are closely correlated and occur simultaneously.

The in vivo signals that affect the commitment decision made by myoblasts during fetal development and myofiber formation have not been identified. One class of protein growth factors, the fibroblast growth factor (FGF), has been shown to be mitogenic for myoblasts in culture and can reduce the tendency to differentiate (Allen et al., 1984; Gospodarowicz et al., 1976, Link-hart et al., 1981). A second growth factor, transforming growth factor beta (TGF-b), is a very potent inhibitor of myoblast differentiation and could be responsible for regulating myogenic cell activities in vivo (Florini et al., 1986). In contrast to the two inhibitors of differentiation, the insulin-like growth factors have been reported to stimulate myoblast proliferation and differentiation in culture (Ewton and Florini, 1980, 1981). The means by which these two antagonistic processes can be stimulated by the same hormone, however, has not been completely clarified. In general, the activities of the growth factors and hormones in embryonic muscle development have yet to be verified in vivo.

Although the specific regulatory agents involved in stimulating differentiation have not been thoroughly documented, many of the gene transitions that occur in association with the terminal step in muscle cell differentiation have been reported (Young and Allen, 1979). From the standpoint of gene regulation, some of the interesting events center around the contractile proteins. The major myofibrillar proteins are synthesized in a coordinate fashion shortly after fusion (Devlin and Emerson, 1978, 1979). These events seemed relatively straightforward, until it became possible to examine them in greater molecular detail. It now appears that there are a series of subtle transitions in expression of specific skeletal muscle isoforms of individual proteins during the in vivo and in vitro development of muscle (reviewed by Caplan et al., 1983). The actin that is first synthesized after myoblast differentiation is of the alpha isoform, but it is alpha-cardiac actin and not alpha-skeletal actin. The transition from alpha-cardiac to alpha-skeletal actin occurs as the myotube matures (Bains et al., 1984; Paterson and Eldridge, 1984). Similarly, myosin light chains and heavy chains (Bandman et al., 1982; Crow et al., 1983; Gauthier et al., 1982; Lowey et al., 1983; Lyons et al., 1983; Whalen et al., 1978) progress through isoform transitions that include fetal, neonatal, and, finally, adult isoforms of the subunits of these proteins. These transitions occur in vivo and are also recapitulated in regenerating muscle (Marechal et al., 1984). Regulators of this developmental scheme have not been elucidated; however, innervation and load-bearing functions may be involved in the feedback that is responsible for alterations in gene expression (Hoffman et al., 1985; Rubinstein and Kelly, 1978).

The environmental factors that regulate the synthesis of specific isoforms and the rate at which these proteins are accumulated are not specifically known, but the mechanisms will be resolved in the near future because genes for these proteins are being studied in detail (reviewed by Robbins et al., 1986; Young et al., 1986). For example, the regulation of alpha-skeletal actin may depend on the DNA sequence in regions of the gene preceding the 5'-untranslated part of the message-coding region (Bergsma et al., 1986; Hu et al., 1986; Melloul et al., 1984). It has been suggested that "transacting" factors in the cytoplasm of myogenic cells interact with nuclear genes to activate their expression (Chiu and Blau, 1984), but the nature of these factors has not been described. In the ease of myosin, thyroid hormone may be involved in myosin heavy-chain synthesis (Butler-Brown et al., 1986; Gambke and Rubinstein, 1984; Izumo et al., 1986). The chemical mediators of the effect of activity level (Brevet et al., 1976; Hoffman et al., 1985) and neurogenic influences (Rubinstein and Kelly, 1978) remain undefined. Detailed information about the structure of important muscle-specific genes, including identification of regulatory sequences, will open the door to studies that are critical to understanding quantitative aspects of muscle protein synthesis regulation, one of the key problems in animal growth research.

Postnatal Development

Understanding the regulation of postnatal muscle growth requires an appreciation of the cellular events underlying the process. Postnatal muscle growth is frequently considered to be due to muscle fiber hypertrophy, in contrast to prenatal muscle growth. This assumption stems from the documented fact that muscle fiber number does not increase dramatically after birth in most animals; consequently, increases in size must be due to hypertrophy (reviewed by Goldspink, 1972; Swatland, 1976).

Although postnatal muscle growth is often thought of in terms of fiber hypertrophy, and not hyperplasia, proliferation and differentiation of myogenic cells are central to the process of postnatal muscle growth. For example, Winick and Noble (1966) demonstrated an 8.5-fold increase in rat muscle DNA from 21 to 133 days of age, corresponding to an 88 percent increase in muscle DNA. Moreover, the relationship between DNA accretion and muscle growth was more firmly established by the findings of Moss (1968) and Swatland (1977), which demonstrated that muscle fiber diameter in growing chicken and pig muscle, respectively, is directly related to the total number of muscle fiber nuclei. Additional studies supporting these results have been reviewed by Allen et al. (1979) and continue to appear regularly in the literature.

Consistent with the point of view that myogenic cell proliferation is critical to the attainment of maximum muscle mass in livestock are studies involving strains of swine that differ in muscle growth potential (Harbison et al., 1976; Powell and Aberle, 1981) and growth studies in cattle (Trenkle et al., 1978). Of the biochemical parameters evaluated in these experiments, DNA accretion and protein/DNA ratios were most intimately related to muscle growth. In addition, the most rapid period of DNA accretion coincided with the most rapid period of muscle growth. The cumulative evidence presented by these and other studies suggests that most muscle fiber DNA found in mature muscle is accumulated postnatally, and the accretion of DNA in muscle is a key factor in regulating muscle growth.

The idea that muscle fiber number is constant beyond the neonatal period had been accepted for years, as had the notion that nuclei within muscle fibers do not replicate their DNA or divide. However, these observations were clearly inconsistent with the large increases in DNA occurring in postnatal muscle. This is explained by the role of satellite cells, the small mononucleated cells that reside between the sarcolemma and basem*nt membrane of muscle fibers (Mauro, 1961). These cells have the ability to proliferate, differentiate, and fuse into adjacent fibers (Moss and Leblond, 1971), which results in the addition of the satellite cell nucleus to the muscle fiber.

Satellite cells are only discernible at the electron microscope level because they look like normal myonuclei that are located adjacent to the sarcolemma inside the fiber. Satellite cells are evenly distributed across the surface of muscle fibers, except for an increased density around the neuromuscular junction (Gibson and Schultz, 1983; Kelly, 1978a). In normal adult muscle from many species, the cells generally make up only a small fraction of the total nuclei associated with fibers, usually ranging from 2 percent to less than 10 percent (Allbrook et al., 1971; Cardasis and Cooper, 1975; Schultz, 1974; Snow, 1977) and varying from one type of fiber and muscle to another; slow-twitch fibers often have a higher percentage of satellite cells than do fast-twitch fibers (Gibson and Schultz, 1983; Kelly, 1978b). Also, there seems to be a greater percentage present in muscles of very young animals and a smaller percentage in muscles of old animals; this is particularly evident in fast-twitch muscle fibers (Gibson and Schultz, 1983).

The myogenic potential of satellite cells and their ability to synthesize DNA, divide, and fuse into existing fibers was established by Moss and Leblond (1971). Their myogenic properties were further documented by isolating mononucleated cells from minced muscle digests (Bischoff, 1974) or by isolating individual muscle fibers (Bischoff, 1975; Konigsberg et al., 1975) and monitoring the division of mononucleated cells in culture. Not only did these mononucleated cells divide but they eventually fused to form multinucleated myotubes. Myotubes formed by satellite cells in vitro synthesize muscle-specific proteins and spontaneously contract in culture (Allen et al., 1980; Cossu et al., 1980).

Although qualitatively they resemble embryonic myogenic cells, satellite cells may well be a separate type of myogenic cell. Cossu et al. (1980) first noted major differences in the morphology of the two, and Allen et al. (1982) found that myotubes derived from satellite cells were only able to synthesize one-third to one-half as much alpha-actin as myotubes formed from neonatal rat muscle. Cossu et al. (1983, 1985) also demonstrated that satellite cells and embryonic myogenic cells responded differently to a tumor promoter, 12-O-tetra-decanoylphorbol-13-acetate (TPA). TPA did not stimulate division or inhibit differentiation of satellite cells, as it did with myogenic cells of embryonic origin. Therefore, factors that stimulate the proliferation or differentiation of embryonic myogenic cells may or may not have the same effect on satellite cells.

Even though the importance of satellite cells to muscle regeneration and normal growth has been appreciated for some time, details of their regulation are only now beginning to emerge. The stimulatory effect of five different growth factors and hormones and the inhibitory effect of one growth factor on satellite cell proliferation have been documented in vitro (Allen, 1986; Allen et al., 1984; Dodson et al., 1985). Three of these proteins are insulin-like growth factors I and II (IGF-I and IGF-II) and insulin (Dodson et al., 1985). These proteins are members of the same gene family and share high degrees of sequence hom*ology (Klapper et al., 1983; Marquardt and Todaro, 1981; Rinderknecht and Humbel, 1978). Insulin is active only at supraphysiological concentrations, which has been explained in terms of its action as an IGF-I analog. Both IGFs (commonly referred to as somatomedins) stimulate satellite cell proliferation at concentrations well within the physiological range. The significance of the IGFs—particularly IGF-I—lies in their relationship to growth hormone. IGF-I mediates the growth hormone signal at the target cell level. Consequently, in vitro data directly link the action of the IGFs to an authentic target cell in postnatal skeletal muscle.

Two additional growth factors active in promoting satellite cell proliferation are the basic (Allen et al., 1984) and acidic (R. E. Allen, University of Arizona, unpublished data) forms of fibroblast growth factor. Unlike the IGFs, however, the basic form of FGF only stimulates proliferation and actually inhibits differentiation. Unfortunately, the physiological role of FGFs or similar proteins has not been established. FGFs have been isolated from a variety of cells and tissues; brain and pituitary tissue are the two most commonly used sources for purification (Gospodarowicz et al., 1976). It is particularly noteworthy that similar protein fractions have been isolated from skeletal muscle (Kardami et al., 1985) and from peritoneal macrophages (Baird et al., 1985). The observations that this growth factor is not freely circulating but can be found in a variety of cells and tissues make it a reasonable candidate for an autocrine or paracrine hormone. This concept may have particular importance in regulation of skeletal muscle regeneration and work-induced hypertrophy, where a local signaling mechanism would seem to be necessary. Insights into the molecular mechanisms of FGF action are sparse, although receptors have been identified (Olwin and Hauschka, 1986). The possible role of FGF or FGF-like proteins as local signals for myogenic cell proliferation is an interesting concept that should be addressed.

Satellite cell culture systems have also been used to evaluate the response of satellite cells to growth hormone, prolactin, luteinizing hormone, thyroid stimulating hormone, epidermal growth factor, platelet-derived growth factor, and nerve growth factor. None of these proteins had the ability to stimulate satellite cell growth in vitro (Allen et al., 1986).

As mentioned previously, an inhibitor of satellite cell proliferation and differentiation has been identified: transforming growth factor beta (TGF-β). In vitro, very low concentrations of TGF-β (< 0.5 ng/ml) can affect both processes (Allen, 1986). This factor is interesting because it can be found in many cell types and has a variety of effects on their functions. It can be either stimulatory or inhibitory, depending on cell type and the presence of other growth factors (Moses et al., 1985). TGF-β apparently is identical to the differentiation inhibitor described by Evinger-Hodges et al. (1982) and Florini et al. (1986).

In summary, it appears that satellite cell activity can be controlled by several protein hormones/growth factors, and it may be the interplay of these factors that determines the state of the cell (quiescence, proliferation, or differentiation). Nutritional and environmental factors that influence muscle fiber DNA accretion in postnatal muscle may be mediated through one or more of these proteins.

Muscle Fiber Protein Metabolism

Muscle protein metabolism encompasses a broad range of cellular activities, many of which are integral parts of energy metabolism in the whole animal. Most notable among these biochemical processes are the deamination of amino acids and the utilization of the carbon skeletons for energy production; supplying amino acids to the liver for gluconeogenesis is another important function. These aspects of muscle protein metabolism are obviously critical to the physiology of the animal, but they are not necessarily directly related to muscle growth. Consequently, this discussion dwells on two broad growth-related processes in muscle: protein synthesis and protein degradation. The quantitative balance between these two activities determines the net accumulation of protein in muscle.

A fundamental concept that has been widely appreciated only within the past decade or so is the fact that muscle protein is in a constant state of flux. Protein is constantly being degraded. It would not be out of the ordinary, for example, to experience a 5 to 10 percent rate of degradation of protein per day. To maintain muscle mass, the muscle would have to synthesize an amount of protein equivalent to 5 to 10 percent of its protein content on a daily basis. The ramifications of this are enormous when one considers the energetic costs of synthesizing one peptide bond and the total number of peptide bonds that must be degraded and resynthesized per day. It is easy to understand why protein turnover represents a significant factor in the "maintenance" energy requirements of an animal. It is also easy to see how the efficiency of growth or production could be enhanced if protein turnover could be altered in a favorable way.

A number of studies have demonstrated the balance between protein synthesis and degradation in domestic animals, laboratory animals, and humans and have revealed a general trend: In growing animals, synthesis and degradation rates are elevated with synthesis rate exceeding degradation rate; as maturity is approached, both synthesis and degradation rates decrease and ultimately reach a low and equal rate. With only minor variations, these trends have been observed in cattle, chickens, and laboratory animals (Lewis et al., 1984; MacDonald and Swick, 1981; McCarthy et al., 1983; Millward and Waterlow, 1978; Millward et al., 1976).

Certain metabolic hormones influence protein turnover; glucocorticoids, for example, cause muscle atrophy by depressing synthesis and degradation (McGrath and Goldspink, 1982). Synthesis rate is apparently depressed to a greater extent than degradation rate. Insulin, on the other hand, causes net accretion of protein, primarily by affecting synthesis rate (Tischler, 1981), and generally antagonizes the glucocorticoid effect on synthesis and degradation (Tomas et al., 1984). Thyroid hormone, T3, can increase degradation rate, but this modulation tends to follow the rate of synthesis (Millward, 1985), so there is a minimal change in protein accretion. Metabolites such as branched-chain amino acids or the keto acids of these amino acids may also be involved in depressing degradation (Mitch and Clark, 1984; Tischler et al., 1982). The integrated response of muscle to the inter-play of metabolites and metabolic hormones is not completely understood but represents an important feature of muscle protein accretion regulation.

In addition to the homeostatic regulation of protein turnover, relative rates of synthesis and degradation are altered during growth. Thus far, the only growth-related hormones that have been implicated in regulating protein degradation are the insulin-like growth factors, the somatomedins. Most of the work in this area has been conducted in vitro, where potent inhibitory effects have been observed (Ballard et al., 1986; Janeczko and Etlinger, 1984). The involvement is somewhat perplexing, since rapid growth rates in young animals are accompanied by increased rates of degradation, not decreased degradation. This point of contention, however, may be related to the in vitro assay system; the key element in the observation may be the decrease in degradation rate relative to synthesis rate.

Several physiological conditions have been shown to affect the rates of synthesis and degradation in skeletal muscle. Included among these are physical influences such as muscle stretching, which leads to hypertrophy (Goldspink, 1978; Summers et al., 1985). In vitro muscle stretching decreases protein degradation (Baracos and Goldberg, 1985). Inflammation, fever, and burns also have a dramatic effect by accelerating protein turnover (Goldberg et al., 1984); the common denominator in these observations and in the stretch-induced alteration in turnover may be calcium metabolism. In vitro, an influx of calcium into cells increases protein degradation (Silver and Etlinger, 1985). Furthermore, the calcium-induced elevation in degradation is of nonlysosomal origin (Furuno and Goldberg, 1986), as evidenced by the failure of lysosomal protease inhibitors to inhibit this calcium-induced response.

At present, an inadequate mechanistic understanding of the biochemical details of protein synthesis and degradation—especially degradation—is blocking progress in research on the regulation of these processes. Nutritional/physiological experimentation has provided an important descriptive base, but future progress depends on cellular and molecular details. As mentioned previously, new information on the regulation of myofibrillar protein isoform transitions and the structure and regulation of genes encoding these proteins will have a dramatic impact on our view of muscle protein synthesis regulation. Molecular details of the interaction of key hormones or their second messengers with myofibrillar protein genes should be forthcoming within the next decade.

It is traditionally assumed that lysosomal enzymes are responsible for intracellular protein degradation. These proteases are contained in lysosomes and are active at acidic pH. Several different proteases are grouped in this class and called cathepsins. Not all cathepsins are able to cleave peptide bonds in myofibrillar proteins; only cathepsins B1, D, H, and L have been found in muscle and are active on myofibrillar protein substrates (see Goll et al., 1983). A problem with attributing myofibrillar protein degradation in skeletal muscle to catheptic proteases is the fact that myofibrils or myofilaments have not been observed in lysosomal structures in muscle. Nor have lysosome-like organelles been observed in association with myofibrils. In addition, treatment of cells with lysosomal enzyme inhibitors failed to suppress calcium-induced protein degradation (Furuno and Goldberg, 1986). Although it has been possible in some eases to show correlations between lysosomal enzyme activity and protein degradation, the cause-and-effect relationship has not been proved.

A more likely mechanism for explaining myofibrillar protein degradation begins with the action of nonlysosomal cytoplasmic pro-teases that selectively cleave certain myofibrillar proteins, resulting in the disassembly of filaments in the myofibril (Dayton et al., 1975). Individual myofibrillar proteins or fragments of these proteins can then be taken up by lysosomes and degraded to individual amino acids. If such a scheme is accurate, one of the rate-limiting steps in the process would be the initial degradation steps accomplished by nonlysosomal pro-teases. Recent evidence suggests that activation of calcium-induced and injury-induced protein degradation in muscle does not involve a lysosomal mechanism (Furuno and Goldberg, 1986).

A couple of strong candidates have been suggested for this degradation role, the first of which is the calcium-dependent neutral protease described by Dayton et al. (1976). This protease, with a molecular weight of 110,000 daltons, is located inside skeletal muscle cells, as well as many other cell types, and is active at neutral pH. In skeletal muscle cells, it is found in the sarcoplasm and not in lysosomal structures or other intracellular membrane-bound organelles. The specificity of this protease is somewhat limited in that it generally cleaves only one or a few peptide bonds in a protein. In the myofibril, the proteins affected are troponin-T, troponin-I, tropomyosin, C-protein, filamin, desmin, the Z-line structure, and possibly titin (Goll et al., 1983). Many of these proteins have regulatory and structural significance. Note, however, that the primary proteins in the myofibril—actin and myosin—are apparently not hydrolyzed by this protease.

Although the regulatory details of this protease have not been elucidated, it is clear that calcium ions and a free sulfhydryl group are required for activity. It is also accepted that two forms of the protease exist, one that requires millimolar concentrations of calcium and another that only requires micromolar concentrations for activity. These are distinctly different proteins that share a high degree of sequence hom*ology. The active sites of these proteins are similar to those of papain, and the calcium-binding regions are similar to those of calmodulin (Emori et al., 1986). To add to the complexity of the system, an inhibitor of these proteases is also found in skeletal muscle. The physiological regulation of these different forms of the enzyme and inhibitor is not clear, but it may be crucial to an understanding of protein degradation and turnover.

Other soluble proteases may also be important components of the myofibril degradation process. Several alkaline or neutral proteases have the ability to hydrolyze actin or myosin, but most of these are not found in muscle cells. Perhaps the degradative system understood in greatest detail is an ATP-dependent protease system (Hershko and Ciechanover, 1982), which is found in many cells but has been most extensively studied in reticulocytes. This system is composed of a small, heat-stable protein called ubiquitin (because of its presence in a highly conserved form in most cells) that interacts with an activating enzyme in an ATP-dependent process to ultimately form a covalent isopeptide bond between the carboxyl group of the C-terminal glycine residue of ubiquitin and an epsilon amino group of a lysine on the target protein. The covalent attachment of ubiquitin is thought to target the protein for protease attack. The pro-teases responsible are ill-defined, but the end products are peptides and a released ubiquitin that can recycle. This system could be responsible for identifying proteins that were damaged structurally or otherwise in-activated. Other protease systems requiring ATP may also be present in cells, but their characterization is far from complete. The primary problems with proposed roles for these ATP-dependent proteolytic systems in muscle protein degradation are the lack of detailed information about the specificity of these systems for muscle proteins and the presence and location of these systems in muscle cells.

During normal growth and in many metabolic states, rates of synthesis and degradation tend to move in tandem. Even during fasting, degradation is depressed and not increased, presumably to spare protein. These observations suggest that during normal growth, protein synthesis may represent the primary site of regulation, and degradation may follow (Millward, 1985). It is virtually impossible to tie together endocrine and nutritional influences on animal protein degradation and the subcellular events that mediate these effects because of the present gap that exists between our knowledge of the cellular and biochemical mechanisms involved in skeletal muscle protein degradation and the whole animal and tissue level descriptions of the process. This does not eliminate the possibility of targeting degradation as a site for muscle growth regulation, but it makes it difficult to devise strategies to manipulate protein degradation to enhance the efficiency of muscle growth in meat animals.

Strategies for Regulating Muscle Development and Growth in Meat-Producing Animals

Significant research areas that can be layered over the muscle-specific problem are the integration of metabolism during growth and the manner in which tissue growth is coordinated within the animal. These topics are more general and, at face value, more pertinent to altering efficiency of protein accretion and the composition of the product than are the studies of specific cellular and biochemical events in developing and growing muscle. But progress in these areas can only proceed as rapidly as progress toward a mechanistic understanding of muscle growth.

Establishing muscle cellularity, in its broadest sense, involves prenatal fiber development and nuclear accretion during postnatal growth. Fiber development is the result of myogenesis that takes place in the developing embryo or fetus. The final event in this cascade of proliferation and differentiation is the fusion of myoblasts into multinucleated myotubes that mature into fibers. Currently, hormones and growth factors that stimulate and inhibit the proliferative and differentiative events in myogenic cells are being identified, but the factors that regulate the number of fibers that are formed from a given cohort of myoblasts have not been considered experimentally. It may be that innervation plays a key role in establishing fiber number and organization in muscle, since innervation is required to sustain fibers. For many years, it has been accepted that major differences in muscle mass in mature animals can be attributed in large part to differences in fiber number. Consequently, alterations in fiber number during late prenatal life would likely result in differences in muscularity. At present, however, there is probably insufficient mechanistic detail to suggest specific approaches. A critical question in this regard is whether it is advisable to increase muscularity prenatally; in cattle, for example, increased management problems associated with dystocia could offset any advantages due to increased muscle growth potential. In swine or poultry, this problem may not be so acute.

Cellularity could conceivably be altered by nuclear accretion postnatally without reproductive problems. Again, we are beginning to understand more about the activation, proliferation, and differentiation of satellite cells, although specific physiological regulators have not yet been confirmed. Assuming that satellite cell activity could be altered and nuclear accretion in fibers could be influenced, satellite cells may be more receptive to manipulation efforts during certain periods of growth than others. Early postnatal growth is the time of greatest satellite cell activity and would correspond to the period of greatest sensitivity to hormones and growth factors. On the other hand, later phases of growth are marked by decreasing nuclear accretion rate; therefore, stimulating additional satellite cell proliferation and differentiation could result in an extension of the rapid muscle growth phase that is normally associated with muscle growth in younger animals.

Affecting changes in muscle growth by altering protein metabolism has been the most commonly considered avenue, primarily because of the erroneous assumption that cellularity does not change after birth. During normal growth, synthesis and degradation tend to move in parallel, with synthesis rate exceeding degradation rate. Consequently, accelerated growth rate is accompanied by an accelerated degradation rate; hence, there is no increase in efficiency of protein accretion. Because these processes seem to be coupled, Millward (1985) suggested that manipulating synthesis may be the most reasonable way to affect protein accretion. Specific alterations in synthesis await increased knowledge of the mechanisms of muscle protein gene regulation and the elucidation of hormones or other external signals, such as electrical stimulation or stretch, that modulate the expression of these genes.

Likewise, strategies for manipulating degradation rate in muscle will not progress beyond the empirical stage without a mechanistic understanding of the proteases involved and their regulation. Protein degradation is, however, an attractive target for postnatal growth manipulation. If degradation rate could be decreased, net rate of protein accretion would be accelerated and less energy would be expended on resynthesizing degraded protein.

To illustrate that the present level of cellular and molecular understanding of important regulatory events is grossly inadequate and, indeed, limiting, consider some current growth-manipulating techniques. Take three growth-altering treatments: growth hormone (GH), steroid hormones and their analogs, and beta-adrenergic agonists. With all three, scientists are still dependent on information that is often one or two decades old, or on empirical observations, the biology of which is still not fully understood. In these cases current biology is not leading the way to new applications; rather, new applications are leading basic biological investigation.

Let us begin with GH. Direct administration of GH to domestic meat animals was first reported in pigs by Truman and Andrews (1955), Henricson and Ullberg (1960), and Machlin (1972). Later, Chung et al. (1985) also reported direct administration of GH to pigs. Dramatic increases in muscle growth in GH-treated pigs (Etherton et al., 1986) could be the result of action at several sites, such as adipose tissue, where GH could be having an antilipogenic effect. If energy is not stored in adipose tissue, it may be more available for growth. It is also possible that when growth processes are stimulated, they demand more energy than do adipose tissue triglyceride storage activities. GH could also be having part of its effect by stimulating higher levels of somatomedins that are, in turn, stimulating satellite cells. Arguments can be made for increased muscle growth as a result of nuclear accretion and subsequent protein accumulation directed by new nuclei. Another plausible alternative may be somatomedinmediated depression of muscle protein degradation. Or, the net effect could be due to a combination of the above. The point is that it is application that is leading scientists to undertake basic biological research.

Next consider the steroid hormones and their analogs. Studies that were mostly empirical in nature gave us diethylstilbestrol. Its application has come and gone from agriculture, yet we still do not know its precise mode of action. Even the action of testosterone on muscle growth is unclear. Trenbolone acetate (TBA) is another ex-ample—it stimulates growth, but again, the mechanism is unknown. In terms of the biological events responsible for muscle growth, it must either directly or indirectly stimulate nuclear accretion, stimulate protein synthesis, or decrease protein degradation. At least one report suggests that protein synthesis is depressed but that protein degradation is depressed to a greater extent, thus leading to a net increase in rate of protein accretion as well as in efficiency of protein gain (Vernon and Buttery, 1976). This in vivo study was not able to address the direct or indirect nature of the action of TBA. In vitro studies are limited; however, TBA does not appear to have a direct effect on protein degradation in L6 muscle cells in culture (Ballard and Francis, 1983).

Another example of application leading basic investigation concerns a class of agents that has received a great deal of attention in recent years, the beta-adrenergic agonists. One of these—clenbuterol—was originally designed as a respiratory drug but was subsequently shown to have a stimulatory effect on rat growth. Since then, it has been used to stimulate growth and feed efficiency in poultry, sheep, and cattle (Baker et al., 1984; Dalrymple et al., 1984; Ricks et al., 1984). A great deal of effort is currently being devoted to understanding how it works. An obvious site of action would be as a lipolytic agent for adipose tissue; however, this alone could not explain the extreme muscle hypertrophy observed in sheep (Beermann et al., 1986). Recently, Kim et al. (1986) reported that the major effect appeared to be on hypertrophy of fast-twitch muscle fibers and that muscle DNA concentration actually decreased in the cimaterol-treated group. Beermann, however, indicated that a significant increase in DNA content was noted in 12-week studies with sheep but that DNA content increased after muscle hypertrophy (D. H. Beermann, personal communication, 1986). In another report, cimaterol was demonstrated to have an inhibitory effect on protein degradation in cultured myotubes from a rat muscle cell line (Forsberg and Merrill, 1986). Evidently, the beta-adrenergic agonists may have multiple sites of action, especially for protein degradation and adipose tissue metabolism, but this conclusion remains highly speculative.

These examples of a few of the most interesting agents currently being investigated for use in stimulating muscle growth not only demonstrate that application is leading investigation, they also provide striking demonstrations that muscle growth in meat animals can be manipulated to increase protein production and decrease triglyceride deposition beyond the normal physiological limits of a particular animal. They also suggest that the muscle growth processes mentioned earlier—protein synthesis and degradation and pre- and post-natal muscle cellularity alterations—repre-sent legitimate targets for growth-regulating strategies.

In the future, several approaches may be used to enhance rate and efficiency of muscle growth, but for now the most promising are administration of recombinant hormones. As indicated, recombinant GH has been shown to have impressive stimulatory effects on growth, feed efficiency, and carcass composition in pigs. Other hormones will surely be investigated in a similar manner. Based on recent research in muscle development, somatomedin-C/IGF-I is a logical choice for such application. New growth factors or combinations of growth factors that affect muscle development, such as fibroblast growth factor and IGF-II, are also candidates.

At present, GH administration entails regular injections during later stages in postnatal life. In a second generation of studies, researchers may wish to effect a permanent change in the cellularity of the animal, such as increased fiber number or myonuclei content. In contrast to approaches that are designed primarily to alter protein metabolism, cellular/developmental changes may only require acute treatments during early, critical stages of development. Therefore, the need for costly, labor-intensive administration schemes could be eliminated, as would potential questions about the presence of drug residues in the final product.

At another level of sophistication, transgenic animals may have a place in livestock production systems. Growth has already been accelerated in transgenic mice carrying a metallothionein-human growth hormone fusion gene (Palmiter et al., 1982) and in mice expressing the metallothionein-human growth hormone releasing factor minigene (Hammer et al., 1985). In addition, these genes have been shown to be transmittable to subsequent generations, although reproduction suffered in some of the initial studies (Hammer et al., 1985). These techniques will undoubtedly be applied in large domestic animals to produce new germplasm. Furthermore, it may be possible to construct and perpetuate the genes of important hormones that can be regulated by coupling the genes to promoters that can be turned on or off at critical periods through nutritional, pharmacological, or environmental manipulation. These approaches will obviously require a more detailed description of the significant regulatory events in muscle growth and the important factors that mediate these events so that appropriate molecular targets can be selected.

Conclusions

Major obstacles exist. New fundamental knowledge of cellular and molecular mechanisms of growth is desperately needed. Technical advances are also needed in the area of delivery systems for effectively administering exogenous agents at specific times and in appropriate amounts. Of critical importance are practical means for targeting the delivery of agents to specific tissues. It is conceivable that a factor could have a beneficial effect on one tissue or organ and a detrimental effect on another. This may be a major impediment to the application of certain hormones or growth factors. Technical advances are still needed in gene transfer and gene construct technology, but progress is occurring rapidly. These are only a few of the problem areas that need to be addressed.

Advances in the production of nutritious muscle protein foods will probably not come by altering the cellular composition of a muscle fiber. Membrane systems and myofibrillar proteins in muscle are highly conserved and may not be amenable to efforts to inflict gross alterations that would provide a more desirable balance of amino acids or reduced cholesterol content. An approach to improving the nutritional attributes of meat products that holds greater promise is one that attempts to reduce the amount of adipose tissue associated with meat products while maintaining palatability. Great advances can be made in the efficient production of muscle protein by providing a growing knowledge base in biology, by rapidly adopting new scientific technologies, and by fostering innovative applied research.

References

  • Allbrook, D. B., M. F. Han, and A. E. Hellmuth. 1971. Population of muscle satellite cells in relation to age and mitotic activity. Pathology 3:233-243. [PubMed: 4107201]

  • Allen, R. E.1986. Transforming growth factor-beta inhibits the IGF-I-induced proliferation and differentiation of skeletal muscle satellite cells. J. Cell Biol. 103(5):120a.

  • Allen, R. E., R. A. Merkel, and R. B. Young. 1979. Cellular aspects of muscle growth. Myogenic cell proliferation. J. Anim. Sci. 49(1):115-127. [PubMed: 500507]

  • Allen, R. E., P. K. McAllister, and K. C. Masak. 1980. Myogenic potential of satellite cells m skeletal muscle of old rats. Mech. Age. Dev. 13:105-109. [PubMed: 7431999]

  • Allen, R. E., P. K. McAllister, K. C. Masak, and G. R. Anderson. 1982. Influence of age on accumulation of α-actin in satellite-cell-derived myotubes in vitro. Mech Age. Dev 18:89-95. [PubMed: 7062790]

  • Alien, R. E., M. V. Dodson, and L. S. Luiten1984. Regulation of skeletal muscle satellite cell proliferation by bovine pituitary fibroblast growth factor. Exp. Cell Res. 152:154-160. [PubMed: 6714317]

  • Allen, R. E., M. V. Dodson, L. K. Boxhorn, S. L. Davis, and K. L. Hossner. 1986. Satellite cell proliferation in response to pituitary hormones. J. Anita. Sci. 62:1596-1601. [PubMed: 3733558]

  • Ashmore, C. R., D. W. Robinson, P. Rattray, and L. Doerr. 1972. Biphasic development of muscle fibers in the fetal lamb. Exp. Neurol. 37(2):241-255. [PubMed: 4118074]

  • Bains, W., P. Ponte, H. Blau, and L. Kedes. 1984. Cardiac actin is the major actin gene product in skeletal muscle cell differentiation in vitro. Mol. Cell. Biol. 4:1449-1453. [PMC free article: PMC368933] [PubMed: 6493226]

  • Baird, A., P. Mormede, and P. Bohlen. 1985. Immunoreactive fibroblast growth factor in cells of peritoneal exudate suggests its identity with macrophage-derived growth factor. Biochem. Biophys. Res. Commun. 126:358-364. [PubMed: 3970698]

  • Baker, P. K., R. H. Dalrymple, D. L. Ingle, and C. A. Ricks. 1984. Use of a β-adrenergic agonist to alter muscle and fat deposition in lambs, J. Anim. Sci. 59:1256-1261.

  • Ballard, F. J., and G. L. Francis. 1983. Effects of anabolic agents on protein breakdown in L6 myoblasts. Biochem. J. 210:243-249. [PMC free article: PMC1154211] [PubMed: 6342615]

  • Ballard, F. J., L. C. Read, G. L. Francis, C. J. Bagley, and J. C. Wallace. 1986. Binding properties and biological potencies of insulin-like growth factors in L6 myoblasts. Biochem. J. 233:223-230. [PMC free article: PMC1153007] [PubMed: 3954726]

  • Bandman, E., R. Matsuda, and R. C. Strohman. 1982. Developmental appearance of myosin heavy and light chain isoforms in vivo and in vitro in chicken skeletal muscle. Der. Biol. 93:508-518. [PubMed: 7141112]

  • Baracos, V. E., and A. L. Goldberg. 1985. Ca2+, interleukin-1 and failure to maintain normal length stimulate protein degradation in isolated skeletal muscle Pp. 603-605 in Intracellular Protein Catabolism, E. A. Khaviallah, editor; , J. S. Bond, editor; , and J. W. C. Bird, editor. (eds.). New York: Alan R. Liss. [PubMed: 3875865]

  • Beermann, D. H., D. E. Hogue, V. K. Fishell, R. H. Dalrymple, and C. A. Ricks. 1986. Effects of cimaterol and fishmeal on performance, carcass characteristics and skeletal muscle growth m lambs, J. Anim. Sci. 62:370-380.

  • Bergsma, D. J., J. M. Grichnik, L. M. A. Gossett, and R. J. Schwartz1986. Delimitation and characterization of cis-acting DNA sequences required for the regulated expression and transcriptional control of the chicken skeletal a-actin gene. Mol, Cell. Biol. 6:2462-2475. [PMC free article: PMC367800] [PubMed: 3785201]

  • Bischoff, R.1974. Enzymatic hberation of myogenic cells from adult rat muscle. Anat. Rec. 180:645-662. [PubMed: 4374100]

  • Bischoff, R.1975. Regeneration of single skeletal muscle fibers in vitroAnat Rec 182:215-236. [PubMed: 168794]

  • Bonner, P. H., and T. R. Adams. 1982. Neural induction of chick myoblast differentiation in culture. Dev Biol. 90:175-184. [PubMed: 7060830]

  • Brevet, A., E Pinto, J. Peaco*ck, and F. E. Stockdale. 1976. Myosin synthesis increased by electrical stimulation of skeletal muscle cell cultures. Science 193:1152-1154. [PubMed: 959833]

  • Buckley, P. A., and I. R. Konigsberg. 1974. Myogenic fusion and the duration of the post-mitotic gap(G1 Dev. Biol. 37:193-212. [PubMed: 4823501]

  • Butler-Brown, G. S., G. S. Cambon, C. Laurent, G. Prulier, M. Toutant, S. Watkins, and R. G. Whalen. 1986. Control of myosin isoenzyme transitions by thyroid hormone in the dwarf mouse strain. J. Physiol. 377:93.

  • Caplan, A. I., M Y. Fiszman, and H. M. Eppenberger. 1983. Molecular and cell isoforms during development. Science 221:921-927. [PubMed: 6348946]

  • Cardasis, C. A., and G. W. Cooper. 1975. An analysis of nuclear numbers in individual muscle fibers during differentiation and growth A satellite cell-muscle fiber growth unit. J. Exp. Zool. 191(3):347-358. [PubMed: 1127400]

  • Chiu, C. P., and H. M. Blau. 1984. Reprogramming cell differentiation in the absence of DNA synthesis. Cell 37:879-887. [PubMed: 6744415]

  • Chung, C. S., T. D. Etherton, and J. P. Wiggins. 1985. Stimulation of swine growth by porcine growth hormone. J. Anim. Sci. 60(1):118-130. [PubMed: 3882649]

  • Cossu, G., B. Zani, M. Colletta, M. Bouche, M. Pacifici, and M. Molinaro. 1980. In vitro differentiation of satellite cells isolated from normal and dystrophic mammalian muscles. A comparison with embryonic myogenic cells Cell Differ. 9:357-368. [PubMed: 7438216]

  • Cossu, G., M. Molinaro, and M. Pacifici. 1983. Differential response of satellite cells and embryonic myoblasts to a tumor promoter. Dev. Biol. 98:520-524. [PubMed: 6873465]

  • Cossu, G., P. Cicinelli, C. Fieri, M. Coletta, and M. Monlinaro. 1985. Emergence of TPA-resistant 'Satellite' cells during muscle histogenesis of human limb. Exp. Cell Res. 160:403-411. [PubMed: 3899691]

  • Crow, M. T., P. S. Olson, and F. E. Stockdale. 1983. Myosin light-chain expression during avian muscle development. J. Cell Biol. 96:736-744. [PMC free article: PMC2112415] [PubMed: 6339522]

  • Dalrymple, R. H., C. A. Ricks, P. K. Baker, J. M. Pensak, P. E. Gingher, and D. L. Ingle1984. Use of the beta-agonist clenbuterol to alter carcass composition in poultry. Poultry Sci. 63:2376-2383. [PubMed: 6531325]

  • Dayton, W. R., D. E. Goll, M. H. Stromer, W. J. Reville, M. G. Zeece, and R. M. Robson. 1975. Some properties of a Ca++-activated protease that may be involved in myofibrillar protein turnover. Pp. 551-577 in Proteases and Biological Control, E. Reich, editor; , D. B. Rifkin, editor; , and E. Shaw, editor. (eds.). Cold Spring Harbor, N.Y.: Cold Spring Harbor Laboratory.

  • Dayton, W. R., D. E. Goll, M. G. Zeece, R. M. Robson, and W. J. Reville. 1976. A Ca2+-activated protease possibly involved in myofibrillar protein turnover. Purification from porcine muscle. Biochemistry 15:2150-2158. [PubMed: 1276130]

  • Devlin, R. B., and C. P. Emerson, Jr.1978. Coordinate regulation of contractile protein synthesis during myoblast differentiation. Cell 13:599-611. [PubMed: 657269]

  • Devlin, R. B., and C. P. Emerson, Jr.1979. Coordinate accumulation of contractile protein mRNAs during myoblast differentiation Dev. Biol. 69:202-216. [PubMed: 446892]

  • Devlin, B. H., and I. R. Konigsberg. 1983. Reentry into the cell cycle of differentiated skeletal myocytes. Dev. Biol. 95:175-192. [PubMed: 6825923]

  • Dodson, M. V., R. E. Allen, and K. L. Hossner. 1985. Ovine somatomedin, multiplication-stimulating activity, and insulin promote skeletal muscle satellite cell proliferation in vitro. Endocrinology 117:2357-2363. [PubMed: 3905359]

  • Eichhorn, J. M., L. J. Coleman, E. J. Wakayama, G. J. Blomquist, D. M. Bailey, and T. G. Jenkins. 1986. Effects of breed type and restricted versus ad libitum feeding on fatty acid composition and cholesterol content of muscle and adipose tissue from mature bovine females. J. Anim. Sci. 63:781-794. [PubMed: 3759706]

  • Emori, Y., H. Kawasaki, H. Sugihara, S. Imajoh, S. Kawashima, and K. Suzuki. 1986. Isolation and sequence analyses of cDNA clones for the large subunits of two isozymes of rabbit calcium-dependent protease. J. Biol. Chem. 261:9465-9471. [PubMed: 2424911]

  • Etherton, T. D., C. M. Evock, C. S. Chung, P. E. Walton, M. N. Sillence, K. A. Magri, and R. E. Ivy. 1986. Stimulation of pig growth performance by long-term treatment with pituitary porcine growth hormone (pGH) and a recombinant pGH. J. Anim. Sci. 63:219 (Abstr.). [PubMed: 3209502]

  • Evinger-Hodges, M. J., D. Z. Ewton, S. C. Seifert, and J. R. Florini. 1982. Inhibition of myoblast differentiation in vitro by a protein isolated from liver cell medium. J. Cell Biol. 93:395-401. [PMC free article: PMC2112867] [PubMed: 7047537]

  • Ewton, D. Z., and J. R. Florini. 1980. Relative effects of the somatomedins, multiplication-stimulating activity, and growth hormone on myoblasts and myotubes in culture. Endocrinology 106:577-583. [PubMed: 7353527]

  • Ewton, D. Z., and J. R. Florini. 1981. Effects of the somatomedins and insulin on myoblast differentiation in vitro. Dev. Biol. 86:31-39. [PubMed: 6169566]

  • Florini, J. R., A. B. Roberts, D. Z. Ewton, S. L. Falen, K. C. Flanders, and M. B. Sp*rn. 1986. Transforming growth factor-b, A very potent inhibitor of myoblast differentiation, identical to the differentiation inhibitor secreted by buffalo rat liver cells. J. Biol. Chem. 261:16509-16513. [PubMed: 3465726]

  • Forsberg, N. E., and G. Merrill. 1986. Effects of cimaterol on protein synthesis and degradation in monolayer culture of rat and mouse myoblasts. J. Anim. Sci. 63:222 (Abstr.).

  • Furuno, K., and A. L. Goldberg. 1986. The activation of protein degradation in muscle by Ca2+ or muscle injury does not involve a lysosomal mechanism. Biochem. J. 237:859-864. [PMC free article: PMC1147067] [PubMed: 3099758]

  • Gambke, B., and N. A. Rubinstein. 1984. A monoclonal antibody to the embryonic myosin heavy chain of rat skeletal muscleJ. Biol. Chem. 259:12092-12100. [PubMed: 6384219]

  • Gauthier, G. F., S. Lowey, P. A. Benfield, and A. W. Hobbs. 1982. Distribution and properties of myosin isozymes in developing avian and mammalian skeletal muscle fibers. J. Cell Biol. 92:471-484. [PMC free article: PMC2112058] [PubMed: 6174531]

  • Gibson, M. C., and E. Schultz. 1983. Age-related differences in absolute numbers of skeletal muscle satellite cells. Muscle Nerve 6:574-580. [PubMed: 6646160]

  • Goldberg, A. L., V. Baracos, P. Rodemann, L. War-man, and C. Dinarello. 1984. Control of protein degradation in muscle by prostaglandins, Ca2+, and leukocytic pyrogen (interleukin 1). Proc. Fed. Am. Soc. Exp. Biol. 43:1301-1306. [PubMed: 6323220]

  • Goldspink, G.1972. Postembryonic growth and differentiation of striated muscle. P. 179 in The Structure and Function of Muscle, Vol. 1, 2nd Ed., G. H. Bourne, editor. (ed.). New York: Academic Press.

  • Goldspink, D. F.1978. The influence of passive stretch on the growth and protein turnover of the denervated extensor digitorum longus muscleBiochem. J. 174:595-602. [PMC free article: PMC1185952] [PubMed: 708412]

  • Goll, D. E., Y. Otsuka, P. A. Nagainis, J. D. Shannon, S. K. Sathe, and M. Muguruma. 1983. Role of muscle proteinases in maintenance of muscle integrity and mass. J. Food Biochem. 7:137-177.

  • Goll, D. E., R. M. Robson, and M. H. Stromer. 1984. Skeletal muscle, nervous system, temperature regulation, and special senses. Pp. 548-580 in Dukes' Physiology of Domestic Animals, 10th Ed., M. J. Swenson, editor. (ed.). Ithaca, N.Y.: Cornell University Press.

  • Gospodarowicz, D., J. Weseman, J. S. Moran, and J. Lindstrom. 1976. Effect of fibroblast growth factor on the division and fusion of bovine myoblasts. J. Cell Biol. 70:395-405. [PMC free article: PMC2109836] [PubMed: 945805]

  • Hammer, R. E., R. L. Brinster, and R. D. Palmiter. 1985. Use of gene transfer to increase animal growth. Cold Spring Harbor Symp. Quant. Biol. 50:379-387. [PubMed: 3006997]

  • Harbison, S. A., D. E. Goll, F. C. Parrish, V. Wang, and E. A. Kline. 1976. Muscle growth in two genetically different lines of swine. Growth 40:253-283. [PubMed: 976769]

  • Hauschka, S. D.1974. Clonal analysis of vertebrate myogenesis. III. Developmental changes in the muscle-colony-forming cells of the human fetal limb. Dev. Biol 37:345-368. [PubMed: 4826281]

  • Henricson, B., and S. Ullberg. 1960. Effects of pig growth hormone on pigs. J. Anim. Sci. 19:1002.

  • Hershko, A., and A. Ciechanover. 1982. Mechanisms of intracellular protein breakdown. Annu. Rev. Biochem. 51:335-364. [PubMed: 6287917]

  • Hoffman, R. K., B. Gambke, L. W. Stephenson, and N. A. Rubinstein. 1985. Myosin transitions in chronic stimulation do not involve embryonic isozymes. Muscle Nerve 8:796-805. [PubMed: 4079958]

  • Holtzer, H., and R. Bischoff. 1970. Mitosis and my-ogenesis. Pp. 29-51 in Physiology and Biochemistry of Muscle as a Food, Vol. 2, E. J. Briskey, editor; , R. G. Cassens, editor; , and B. B. Marsh, editor. (eds.). Madison:University of Wisconsin Press.

  • Hu, M. C. T., S. B. Sharp, and N. Davidson. 1986. The complete sequence of the mouse skeletal α-actin gene reveals several conserved and inverted repeat sequences outside of the protein-coding region. Mol. Cell. Biol. 6:15-25. [PMC free article: PMC367479] [PubMed: 3023820]

  • Izumo, S., B. Nadal-Ginard, and V. Mahdavi. 1986. All members of the MHC multigene family respond to thyroid hormone in a highly tissue-specific manner. Science 231:597-600. [PubMed: 3945800]

  • Janeczko, R. A., and J. D. Etlinger. 1984. Inhibition of intracellular proteolysis in muscle cultures by multiplication-stimulating activity. Comparison of effects of multiplication-stimulating activity and insulin on proteolysis, protein synthesis, amino acid uptake, and sugar transport, J. Biol, Chem. 259:6292-6297. [PubMed: 6373754]

  • Kardami, E., D. Spector, and R. C. Strohman. 1985. Myogenic growth factor present in skeletal muscle is purified by heparin-affinity chromatography. Proc. Natl. Acad. Sci. USA 82:8044-8047. [PMC free article: PMC391438] [PubMed: 3865214]

  • Kelly, A. M.1978. a. Perisynaptic satellite cells in the developing and mature rat soleus muscle. Anat. Rec. 190:891-904. [PubMed: 637326]

  • Kelly, A. M.1978. b. Satellite cells and myofiber growth in the rat soleus and extensor digitorum longus muscles. Dev. Biol. 65:1-10. [PubMed: 680348]

  • Kelly, A. M., and S. I. Zacks. 1969. The histogenesis of rat intercostal muscle. J. Cell Biol. 42:135-153. [PMC free article: PMC2107573] [PubMed: 5786979]

  • Kim, Y. S., Y. B. Lee, C. R. Ashmore, and R. H. Dalrymple. 1986. Effect of the repartitioning agent, cimaterol (CL 263,780), on growth, carcass characteristics and skeletal muscle cellularity of lambs. J. Anim. Sci. 63:221 (Abstr.).

  • Klapper, D. G., M. E. Svoboda, and J. J. Van Wyk. 1983. Sequence analysis of somatomedin-C. Confirmation of identity with insulin-like growth factor I. Endocrinology 112:2215-2217. [PubMed: 6343062]

  • Konigsberg, U. R., B. H. Lipton, and I. R. Konigsberg. 1975. The regenerative response of single mature muscle fibers. Dev. Biol. 45:260-275. [PubMed: 1193298]

  • Lewis, S. E. M., F. J. Kelly, and D. F. Goldspink. 1984. Pre- and post-natal growth and protein turnover in smooth muscle, heart and slow- and fast-twitch skeletal muscles of the rat. Biochem J. 217:517-526. [PMC free article: PMC1153244] [PubMed: 6199021]

  • Linkhart, T. A., C. H. Clegg, and S. D. Hauschka. 1981. Myogenic differentiation in permanent clonal mouse myoblast cell lines: Regulation by macro-molecular growth factors in the culture medium Dev.Biol. 86:19-30. [PubMed: 7286393]

  • Lowey, S., P. A. Benfield, D. D. LeBlanc, and G. S. Waller. 1983. Myosin isozymes in avian skeletal muscles. I. Sequential expression of myosin isozymes in developing chicken pectoralis muscles. J. Mus. Res. Cell Motil. 4:695-716. [PubMed: 6230370]

  • Lyons, G. E., J. Haselgrove, A. M. Kelly, and N. A. Rubinstein1983. Myosin transitions in developing fast and slow muscles of the rat hindlimb. Differentiation 25:168-175. [PubMed: 6363184]

  • MacDonald, M. L., and R. W. Swick. 1981. The effect of protein depletion and repletion on muscle-protein turnover in the chick. Biochem. J. 194:811-819. [PMC free article: PMC1162817] [PubMed: 6171262]

  • Machlin, L. J.1972. Effect of porcine growth hormone on growth and carcass composition of the pig. J. Anim. Sci. 35(4):794. [PubMed: 4538590]

  • Marechal, G., K. Schwartz, G. Beckers-Bleukx, and E. Ghins. 1984. Isozymes of myosin in growing and regenerating rat muscles. Eur J. Biochem. 138:421-428. [PubMed: 6697997]

  • Marquardt, H., and G. J. Todaro. 1981. Purification and primary structure of a polypeptide with multiplication-stimulating activity from rat liver cell cultures. hom*ology with human insulin-like growth factor II. J Biol. Chem. 256:6859-6865. [PubMed: 7016879]

  • Mauro, A.1961. Satellite cell of skeletal muscle fibers. J. Biophys. Biochem. Cytol. 9:493-495. [PMC free article: PMC2225012] [PubMed: 13768451]

  • McCarthy, F. D., W. G. Bergen, and D. R. Hawkins. 1983. Muscle protein turnover in cattle of differing genetic backgrounds as measured by urinary Nt-methylhistidine excretion. J. Nutr. 113:2455-2463. [PubMed: 6655509]

  • McGrath, J. A., and D. F. Goldspink. 1982. Glucocorticoid action on protein synthesis and protein breakdown in isolated skeletal muscles. Biochem. J. 206:641-645. [PMC free article: PMC1158634] [PubMed: 7150268]

  • McLennan, I. S.1983. Neural dependence and independence of myotube production in chicken hind-limb muscles. Dev. Biol. 98:287-294. [PubMed: 6873458]

  • Melloul, D., B. Aloni, J. Calvo, D. Yaffe, and U. Nudel. 1984. Developmentally regulated expression of chimeric genes containing muscle actin DNA sequences in transfected myogenic cells. Eur. Mol. Biol. Org. J. 3:983-990. [PMC free article: PMC557461] [PubMed: 6329749]

  • Miller, J. B., and F. E. Stockdale. 1986. Developmental origins of skeletal muscle fibers. Clonal analysis of myogenic cell lineages based on expression of fast and slow myosin heavy chains. Proc. Natl. Acad. Sci. USA 83:3860-3864. [PMC free article: PMC323624] [PubMed: 3520558]

  • Millward, D. J.1985. The physiological regulation of proteolysis in muscle. Biochem. Soc. Trans. 13:1023-1026. [PubMed: 3879228]

  • Millward, D. J., and J. C. Waterlow. 1978. Effect of nutrition on protein turnover in skeletal muscle. Fed. Proc. 37:2283-2290. [PubMed: 350634]

  • Millward, D. J., P. J. Garlick, D. O. Nnanyelugo, and J. C. Waterlow. 1976. The relative importance of muscle protein synthesis and breakdown in the regulation of muscle mass. Biochem. J. 156:185-188. [PMC free article: PMC1163731] [PubMed: 133677]

  • Mitch, W. E., and A. S. Clark. 1984. Specificity of the effects of leucine and its metabolites on protein degradation in skeletal muscle. Biochem. J. 222:579-586. [PMC free article: PMC1144218] [PubMed: 6487265]

  • Moses, H. L., R. F. Tucker, E. B. Leof, R. J. Coffey, Jr., J. Halper, and G. D. Shipley. 1985. Type-β transforming growth factor is a growth stimulator and a growth inhibitor. Pp. 65-71 in Cancer Cells 3/Growth Factors and Transformation, J. Feramisco, editor; , B. Ozanne, editor; , and C. Stiles, editor. (eds.). Cold Spring Harbor, N.Y.: Cold Spring Harbor Laboratory.

  • Moss, F. P.1968. The relationship between the dimensions of the fibres and the number of nuclei during normal growth of skeletal muscle in the domestic fowl. Am. J. Anat. 122:555-564. [PubMed: 5691188]

  • Moss, F. P., and C. P. Leblond. 1971. Satellite cells as the source of nuclei in muscles of growing rats. Anat. Rec. 170:421-436. [PubMed: 5118594]

  • Nadal-Ginard, B.1978. Commitment, fusion and biochemical differentiation of a myogenic cell line in the absence of DNA synthesis. Cell 15:855-864. [PubMed: 728992]

  • Nadal-Ginard, B., H. T. Nguyen, and V. Mahdavi. 1984. Induction of muscle-specific genes in the absence of commitment is reversible. Exp. Biol. Med. 9:250-259.

  • Nguyen, G. T., R. M. Medford, and B. Nadal-Ginard. 1983. Reversibility of muscle differentiation in the absence of commitment. Analysis of a myogenic cell line temperature-sensitive for commitment. Cell 34:281-293. [PubMed: 6683997]

  • Olwin, B. B., and S. D. Hauschka. 1986. Identification of the fibroblast growth factor receptor of Swiss 3T3 cells and mouse skeletal muscle myoblasts. Biochemistry 25:3487-3492. [PubMed: 3013291]

  • Palmiter, R. D., R. L. Brinster, R. E. Hammer, M. E. Trumbauer, M. G. Rosenfeld, N. C. Birnberg, and R. M. Evans. 1982. Dramatic growth of mice that develop from eggs microinjected with metallothionein-growth hormone fusion genes. Nature 300.611-614. [PMC free article: PMC4881848] [PubMed: 6958982]

  • Paterson, B. M., and J. D. Eldridge. 1984. α-Cardiac actin is the major sarcomeric isoform expressed in embryonic avian skeletal muscle. Science 224:1436-1438. [PubMed: 6729461]

  • Peter, J. B., R. J. Barnard, V. R. Edgerton, C. A. Gillespie, and K. E. Stempel. 1972. Metabolic profiles of three fiber types of skeletal muscle in guinea pigs and rabbits. Biochemistry 11:2627-2633. [PubMed: 4261555]

  • Powell, S. E., and E. D. Aberle. 1981. Skeletal muscle and adipose tissue cellularity in runt and normal birth weight swine. J. Anim. Sci. 52:748-756. [PubMed: 7196399]

  • Quinn, L. S., M. Nameroff, and H. Holtzer. 1984. Age-dependent changes in myogenic precursor cell compartment sizes. Exp. Cell Res. 154:65-82. [PubMed: 6468535]

  • Reiser, R.1975. Fat has less cholesterol than lean. J. Nutr. 105:15-16.

  • Rhee, K. S., T. R. Dutson, and G. C. Smith. 1982. Effect of changes in intermuscular and subcutaneous fat levels on cholesterol content of raw and cooked beef steaks. J. Food Sci. 47:1638-1642.

  • Ricks, C. A., R. H. Dalrymple, P. K. Baker, and D. L. Ingle. 1984. Use of a β-agonist to alter fat and muscle deposition in steers. J. Anim. Sci. 59:1247-1255.

  • Rinderknecht, E., and R. E. Humbel. 1978. Primary structure of human insulin-like growth factor II. FEBS 89:283-286. [PubMed: 658418]

  • Robbins, J., T. Horan, J. Gulick, and K. Kropp. 1986. The chicken myosin heavy chain family. J. Biol. Chem. 261:6606-6612. [PubMed: 3009464]

  • Rubinstein, N. A., and A. M. Kelly. 1978. Myogenic and neurogenic contributions to the development of fast and slow twitch muscles in the rat. Dev. Biol. 62:473-485. [PubMed: 146629]

  • Rutz, R., and S. Hauschka. 1982. Clonal analysis of vertebrate myogenesis. VII. Heritability of muscle colony type through sequential subclonal passages in vitro. Dev. Biol. 91:103-110. [PubMed: 7095254]

  • Savell, J. W., H. R. Cross, and G. C. Smith. 1986. Percentage ether extractable fat and moisture content of beef longissimus muscle as related to USDA marbling score. J. Food Sci. 51:838-840.

  • Schultz, E.1974. A quantitative study of the satellite cell population in postnatal mouse lumbrical muscle. Anat. Rec. 180:589-596. [PubMed: 4440878]

  • Seed, J., and S. D. Hauschka. 1984. Temporal separation of the migration of distinct myogenic precursor populations into the developing chick wing bud. Dev. Biol. 106:389-393. [PubMed: 6500180]

  • Silver, G., and J. D. Etlinger. 1985. Regulation of myofibrillar accumulation in chick muscle cultures. Evidence for the involvement of calcium and lysosomes in non-uniform turnover of contractile proteins. J. Cell Biol. 101:2383-2391. [PMC free article: PMC2114019] [PubMed: 3934180]

  • Snow, M. H.1977. Myogenic cell formation in regenerating rat skeletal muscle injured by mincing. II. An autoradiographic study. Anat. Rec. 188:201-218. [PubMed: 869238]

  • Stockdale, F. E., and H. Holtzer. 1961. DNA synthesis and myogenesis. Exp. Cell Res. 24:508-520. [PubMed: 13917284]

  • Stromer, M. H., D. E. Goll, and J. H. Roberts. 1966. Cholesterol in subcutaneous and intramuscular lipid depots from bovine carcasses of different maturity and fatness. J. Anim. Sci. 25:1145-1147.

  • Summers, P. J., C. R. Ashmore, Y. B. Lee, and S. Ellis. 1985. Stretch-induced growth in chicken wing muscles. Role of soluble growth-promoting factors. J. Cell. Physiol. 125:288-294. [PubMed: 4055912]

  • Swatland, H. J.1976. Recent research on postnatal muscle development in swine. Proc. Recip. Meat Conf. 29:86-103.

  • Swatland, H. J.1977. Accumulation of myofiber nuclei in pigs with normal and arrested development. J. Anim. Sci. 44:759-764. [PubMed: 863797]

  • Tischler, M. E.1981. Hormonal regulation of protein degradation in skeletal and cardiac muscle. Life Sci. 28:2569-2576. [PubMed: 7022077]

  • Tischler, M. E., M. Desautels, and A. L. Goldberg. 1982. Does leucine, leucyl-tRNA, or some metabolite of leucine regulate protein synthesis and degradation in skeletal and cardiac muscle?J. Biol. Chem. 257:1613-1621. [PubMed: 6915936]

  • Tomas, F. M., J. Murray, and L M. Jones. 1984. Interactive effects of insulin and corticosterone on myofibrillar protein turnover in rats as determined by N-methylhistidine excretion. Biochem. J. 220:469-479. [PMC free article: PMC1153649] [PubMed: 6378188]

  • Trenkle, A., D. L. DeWitt, and D. G. Topel. 1978. Influence of age, nutrition and genotype on carcass traits and cellular development of the M. Longissimus of cattle. J. Anim. Sci. 46:1597-1603.

  • Truman, E. J., and F. N. Andrews. 1955. Some effects of purified anterior pituitary growth hormone on swine. J. Anim. Sci. 14:7.

  • Vernon, B. G., and P. J. Buttery. 1976. Protein turnover in rats treated with trienbolone acetate. Br. J. Nutr. 36:575-579. [PubMed: 1009081]

  • Watt, B. K., and A. L. Merrill. 1963. Composition of foods—raw, processed, prepared. In USDA Agricultural Handbook 8. Washington, D.C.: U.S. Department of Agriculture.

  • Whalen, R G., G. S. Butler-Browne, and F. Gros. 1978. Identification of a novel form of myosin light chain present in embryonic muscle tissue and cultured muscle cells. J. Mol. Biol. 126:415-431. [PubMed: 745235]

  • White, N. K., P. H. Bonner, D. R. Nelson, and S. D. Hauschka. 1975. Clonal analysis of vertebrate myogenesis. IV. Medium-dependent classification of colony forming cells. Dev. Biol. 44:346-361. [PubMed: 1132597]

  • Winick, M., and A. Nobel. 1966. Cellular response in rats during malnutrition at various ages. J. Nutr. 89:300-306. [PubMed: 5913937]

  • Young, R. B., and R. E. Allen. 1979. Transitions in gene activity during development of muscle fibers. J. Anim. Sci. 48:837-852. [PubMed: 383671]

  • Young, R. B., D. M. Moriarity, and C. E. McGee. 1986. Structural analysis of myosin genes using recombinant DNA techniques. J. Anim. Sci. 63:259-268. [PubMed: 3525491]

Muscle Cell Growth and Development (2024)
Top Articles
Latest Posts
Article information

Author: Corie Satterfield

Last Updated:

Views: 6593

Rating: 4.1 / 5 (42 voted)

Reviews: 81% of readers found this page helpful

Author information

Name: Corie Satterfield

Birthday: 1992-08-19

Address: 850 Benjamin Bridge, Dickinsonchester, CO 68572-0542

Phone: +26813599986666

Job: Sales Manager

Hobby: Table tennis, Soapmaking, Flower arranging, amateur radio, Rock climbing, scrapbook, Horseback riding

Introduction: My name is Corie Satterfield, I am a fancy, perfect, spotless, quaint, fantastic, funny, lucky person who loves writing and wants to share my knowledge and understanding with you.