You are on page 1of 14

Meat Science 84 (2010) 257–270

Contents lists available at ScienceDirect

Meat Science
journal homepage: www.elsevier.com/locate/meatsci

Review

A second look into fibre typing – Relation to meat quality


L. Lefaucheur *
Institut National de la Recherche Agronomique (INRA), Unité Mixte de Recherches 1079, Systèmes d’Elevage, Nutrition Animale et Humaine, Domaine de la Prise,
F-35590 Saint-Gilles, France

a r t i c l e i n f o a b s t r a c t

Article history: Despite intensive research, a large variation in meat quality is still observed in most meat producing spe-
Received 4 March 2009 cies. It is widely accepted that myofibre type composition is an important source of variation in meat
Received in revised form 16 April 2009 quality. However, the identification of specific and universal relationships between myofibre character-
Accepted 3 May 2009
istics, growth performance and meat quality traits remains a challenge. After the presentation of recent
knowledge underlying fibre typing, this review describes the involvements of Ca2+-dependent mecha-
nisms, and the energy state of the myofibres in the control of contractile and metabolic properties, with
Keywords:
a special attention to the AMP-activated protein kinase pathway and mitochondrial compartment. In
Muscle fibres
Myosin heavy chain
order to identify muscle components which could mask specific relationships between fibre type compo-
Energy metabolism sition and meat quality, an analysis of the interactions between myofibres and other muscle cellular com-
Growth ponents is presented. After a brief description of myogenesis, the significance of the total number of
Meat quality fibres, myofibre cross-sectional area and fibre type composition for growth performance and meat quality
is presented. Then, some genetic and environmental factors are proposed as possible tools to control meat
quality trough the modulation of fibre type characteristics. Finally, a conclusion makes the point on bot-
tlenecks still preventing the identification of specific relationships between fibre characteristics, growth
performance and meat quality, and suggests future perspectives such as direct selection on fibre traits
and study of correlated responses, the development of in vitro approaches using cell cultures, manipula-
tion of myogenesis during the fetal period, and the production and use of genetically modified animals.
Ó 2009 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
2. Fibre type diversity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
2.1. Contractile type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
2.2. Metabolic type. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
2.3. Biological characteristics of fibre types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
2.4. Control of contractile and metabolic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
2.4.1. Ca2+-dependent pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
2.4.2. Energy state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
2.4.3. Mitochondrial respiration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
2.5. Relationships between myofibre type composition and other muscle components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
3. Significance of myofibre type for growth performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
3.1. Total number of fibres and myofibre cross-sectional area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
3.2. Fibre type composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
4. Fibre type and meat quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
4.1. Total number of fibres and myofibre cross-sectional area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
4.2. Fibre type composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
5. Tools to control meat quality through fibre type composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
5.1. Genetic factors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
5.2. Exercise, ambient temperature and nutrition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
6. Conclusion and future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266

* Tel.: +33 2 23 48 56 43; fax: +33 2 23 48 50 80.


E-mail addresses: Louis.Lefaucheur@rennes.inra.fr, louis.lefaucheur@wanadoo.fr

0309-1740/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.meatsci.2009.05.004
258 L. Lefaucheur / Meat Science 84 (2010) 257–270

1. Introduction ities to a number of genetic, nutritional and environmental condi-


tions encountered during the breeding phase of animal
In all species raised for meat production, genetic selection and production, such as physical exercise. After a presentation of re-
improvement of nutrition and breeding conditions has led to a cent data underlying myofibre typing, the present review will fo-
tremendous increase in the efficiency of animal production and cus on the significance of myofibre types for growth performance
carcass composition by decreasing carcass fatness, and increasing and meat quality, and will end with some perspectives dealing
muscle yield. Several studies suggest that such practice would with the control of meat quality through manipulation of fibre
have adversely affected some aspects of lean meat quality (Cam- type composition during the breeding phase.
eron, Nute, Brown, Enser, & Wood, 1999), but the underlying bio-
logical mechanisms remain to be clearly identified. It is widely 2. Fibre type diversity
accepted that myofibre type composition is an important source
of variation in meat quality (Karlsson, Klont, & Fernandez, Myofibres represent 75–90% of the muscle volume and are mul-
1999). However, because of the highly heterogeneous tissue and tinucleated due to their formation by fusion of myoblasts during
cellular composition of skeletal muscle, and the numerous factors fetal development. Their diameter ranges from 10 to 100 lm and
affecting meat quality (pre-, peri-, and post-mortem), it remains they are typically classified according to their contractile and met-
difficult to identify what muscle biological traits are specifically abolic properties.
involved in determination of meat quality. On average, skeletal
muscle contains about 75% water, 19% protein, 0.5–8% lipid and 2.1. Contractile type
1% glycogen, and is composed of several tissues and cell types
including myofibres, along with connective tissue, intramuscular Based on differences in the sensitivity of the acto-myosin ATP-
adipocytes, vascular and nervous tissues. Moreover, myofibres ase activity to pH preincubation, three main fibre types have been
represent a heterogeneous population differing by structural, con- conventionally determined by histochemistry in adult skeletal
tractile, metabolic and physiological properties. The functional muscle, i.e. types I, IIA and IIB fibres (Brooke & Kaiser, 1970). The
diversity between fibre types, muscles and species has been underlying molecular basis of this typology resides in the polymor-
achieved during evolution through Darwinian natural selection. phism of myosin heavy chains (MyHC), each isoform being en-
Myofibres can be characterized by their total number (TNF), coded by a single gene (Weiss et al., 1999). However, four adult
cross-sectional area (CSA), length, and contractile and metabolic MyHC isoforms have been identified in mouse, rat, guinea pig
properties. Skeletal muscle exhibits remarkable adaptive capabil- and rabbit skeletal muscles, i.e. types I, IIa, IIx and IIb isoforms

Fig. 1. Serial sections of longissimus muscle from a Large White (A, C, E) and Meishan (B, D, F) pig at 62 kg BW. Histochemical mATPase staining after preincubation at pH 4.35
(A, B). In situ hybridization using 35S labelled MyHC IIx (C, D) and IIb (E, F) riboprobes. The bar indicates 100 lm. Adapted from Lefaucheur et al. (2004).
L. Lefaucheur / Meat Science 84 (2010) 257–270 259

(Bär & Pette, 1988; Schiaffino et al., 1989), which lead to revise the Thus, a large amount of ATP is hydrolyzed during contraction
conventional classification to take into account what MyHC were through activation of the acto-myosin ATPase, and a harmonious
expressed, i.e. types I, IIa, IIx and/or IIb MyHC. The speed of con- functioning needs a balance between ATP consumption and pro-
traction increases in the rank order I < IIa < IIx < IIb (Schiaffino & duction to prevent ATP depletion. Two major metabolic pathways
Reggiani, 1996). Despite the presence of the IIb MyHC gene in all are used to regenerate ATP in skeletal muscle, i.e. the aerobic or
mammals, the type IIb MyHC isoform was initially found in small oxidative pathway through which glycogen, glucose, amino acids,
mammals, and reported to be unexpressed in skeletal muscles of ketone bodies, and lipids can be oxidized in the mitochondria with
large mammals such as human, monkey, cattle, horse, goat and a high oxygen requirement, and the anaerobic or glycolytic path-
dog (Acevedo & Rivero, 2006; Arguello, LopezFernandez, & Rivero, way through which intrafibre glycogen stores are rapidly con-
2001; Ennion, Sant’Ana Pereira, Sargeant, Young, & Goldspink, verted to lactate with no oxygen requirement. Aerobic
1995; Maccatrozzo, Patruno, Toniolo, Reggiani, & Mascarello, metabolism is sufficient to cover energy demand when oxygen is
2004; Maccatrozzo et al., 2007; Rivero, Serrano, Barrey, Valette, & adequate and muscle is working with low intensity, but if contrac-
Jouglin, 1999; Serrano et al., 2001; Smerdu, Karsch-Mizrachi, tion proceeds faster and with high intensity, oxygen becomes lim-
Campione, Leinwand, & Schiaffino, 1994; Smerdu, Strbenc, Mezna- iting and the muscle will have to use the rapid anearobic glycolytic
ric-Petrusa, & Fazarinc, 2005; Tanabe, Muroya, & Chikuni, 1998; pathway to supply energy. The relative importance of these two
Toniolo et al., 2005, 2007). Until recently, the hypothesis among metabolic pathways has been used to type myofibres as oxidative,
researchers was that type IIb MyHC, the fastest isoform, was only oxido-glycolytic or glycolytic (Ashmore & Doerr, 1971; Peter, Bar-
expressed in small mammals in relation to the relative higher nard, Edgerton, Gillespie, & Stemple, 1972). A mitochondrial en-
speed of contraction of their muscles compared to large species. zyme, succino-dehydrogenase (SDH), is commonly used to type
However, recent data using immunocytochemistry, in situ hybrid- myofibres as red (R, SDH+) or white (W, SDH) by histochemistry.
ization, and real time RT-PCR demonstrated that the IIb MyHC iso- By combining conventional acto-myosin ATPase after preincuba-
form was highly expressed in glycolytic muscles of the pig and tion at pH 4.35 with SDH staining on serial sections, pig longissi-
llama (Da Costa, Blackley, Alzuherri, & Chang, 2002; Graziotti,
Palencia, Delhon, & Rivero, 2004; Graziotti, Rios, & Rivero, 2001;
Lefaucheur et al., 1998; Lefaucheur, Ecolan, Plantard, & Gueguen, 100 100
2002; Lefaucheur, Milan, Ecolan, & Le Callennec, 2004). Therefore,
the expression of the IIb MyHC was not restricted to small mam-
80 80
mals as previously suggested, and the conventional histochemical
fibre typing in types I, IIA and IIB (Brooke & Kaiser, 1970) appeared
to be not well adapted for pig and llama skeletal muscles where 60 60
four MyHC are present. The limitation of the conventional fibre
%

%
typing in the pig is clearly illustrated when comparing longissimus
40 40
muscle fibre type composition between the Large White (LW) and
Meishan breeds (Fig. 1). Indeed, no difference in the proportion of
conventional types I, IIA and IIB fibres was observed between the 20 20
two breeds, even though MyHC IIb and IIx were the prominent iso-
forms in LW and Meishan pigs, respectively (Lefaucheur et al.,
0 0
2004). Recently, expression of the IIb MyHC isoform has been re- s sa a ax x xb b
IIA

W
I

IIB

IIB

ported in longissimus and semitendinosus muscles of few bulls


within the Blonde d’Aquitaine breed (Picard & Cassar-Malek, 2009). Fig. 2. Correspondance between the conventional histochemical fibre type classi-
Myosin heavy chain is the most abundant protein expressed in fication (I, IIA, IIBR and IIBW, left) and the presence of the myosin heavy chains I (s),
skeletal muscle, comprising about 35% of the protein pool. The IIa (a), IIx (x) and IIb (b) revealed by immunocytochemistry (right) in Large White
ATPase activity of MyHC is fundamental in determining the rate pig longissimus muscle at 108 kg BW. Hybrid fibres are noted sa, ax and xb. Black
and white areas denote succino-dehydrogenase positive (oxidative) and negative
of cross-bridge cycling, and thus the speed of contraction. At least
(glycolytic) fibres, respectively. Percentages have been estimated from a total of
eight distinct skeletal MyHC genes are present in mammalian stri- 13.000 myofibres using computerized image analysis. Lefaucheur et al.
ated muscles, including two developmental (embryonic and fetal), (unpublished).
one slow (type I or b-cardiac), one a-cardiac, three adult fast type II
MyHC (IIa, IIx, and IIb), and the extraocular MyHC, a specialized
isoform. These genes are located in two clusters in all mammals Table 1
Biological characteristics of individual fibre typesa. Adapted from Bacou and Vigneron
studied so far (Davoli et al., 2002; Mahdavi, Chambers, & Nadal-
(1988), Essén-Gustavsson et al. (1994), Pette and Staron (2000), Quiroz-Rothe and
Ginard, 1984; Shrager et al., 2000; Weiss et al., 1999; Weiss, Rivero (2004), Schiaffino and Reggiani (1996).
Schiaffino, & Leinwand, 1999). The skeletal MyHC gene cluster con-
I IIa IIx IIb
tains the embryonic, IIa, IIx, IIb, fetal and extraocular MyHC in this
order, whereas the cardiac cluster consists in the b MyHC (slow Contraction speed + +++ ++++ +++++
type I) upstream of the a-cardiac MyHC gene. Expression of MyHC Myofibrillar ATPase + +++ ++++ +++++
Oxidative metabolism +++++ ++++, +++++ +, ++ +
is reported to be transcriptionally regulated (Cox & Buckingham,
Glycolytic metabolism + ++++ ++++ +++++
1992). However, very recent data suggest that MyHC gene regula- Hexokinase +++++ +++ + +
tion would be more complex than previously appreciated due to GLUT-4 +++++ +++ + +
the presence of bidirectional intergenic promoters coordinating Phosphocreatine + +++++ +++++ +++++
Glycogen + +++++ ++++ +++++
the expression of the different MyHC (Haddad et al., 2008; Rinaldi
Triglycérides +++++ ++ + +
et al., 2008). Vascularization +++++ +++ +, ++ +
Myoglobine +++++ ++++ ++ +
2.2. Metabolic type Buffering capacity + ++++ +++++ +++++
Diameter ++ +, ++ ++++ +++++
Fatigue resistance +++++ ++++ ++ +
Skeletal muscle is characterized by a dramatic increase in en-
a
ergy consumption and turnover in response to physical exercise. +, very low; ++, low; +++, medium; ++++, high; +++++, very high.
260 L. Lefaucheur / Meat Science 84 (2010) 257–270

mus muscle has been shown to contain 9.5% type I, 6.5% type IIA, content and insulin sensitivity. Therefore, the level of intramyofi-
10% type IIBR (SDH+), and 74% type IIBW (SDH) (Larzul et al., bre TAG does not seem to be a causal factor inducing insulin resis-
1997). The correspondence between this classification and the tance which would rather be mediated by intermediate lipid
presence of the MyHC I (s), IIa (a), IIx (x) and IIb (b) revealed by metabolites, such as diacylglycerol and/or ceramide (Turinsky,
immunocytochemistry shows that fibres containing either IIx, O’Sullivan, & Bayly, 1990).
(IIx + IIb) (hybrid fibres), or only IIb MyHC are mostly IIBW fibres
(Fig. 2), denoting that conventional histochemical fibre typing in 2.4. Control of contractile and metabolic properties
types I, IIA, IIBR and IIBW cannot fully distinguish fibres containing
different MyHC, in particular IIx and IIb MyHC. 2.4.1. Ca2+-dependent pathways
In normal cells, basal sarcoplasmic Ca2+ level is maintained
2.3. Biological characteristics of fibre types close to 107 M, which is about 104-fold lower than in SR. During
contraction, membrane depolarization stimulates the release of
Individual fibre types exhibit different contractile, metabolic, Ca2+ from the SR, which increases sarcoplasmic Ca2+ concentration
physiological, chemical and morphological characteristics as pre- up to 100-fold. The increase in sarcoplasmic Ca2+, in turn, serves as
sented in Table 1. Slow-twitch oxidative type I fibres exhibit low a second messenger through interaction with troponin C to trigger
myofibrillar and sarcoplasmic reticulum (SR) ATPase activities, contraction, and with calcineurin, a Ca2+/calmodulin-activated ser-
and can sustain prolonged low power work in association with a ine-threonine phosphatase, Ca2+-calmodulin-dependent protein
well-developed oxidative metabolism, an efficient vascularization, kinases (CaMK), and Ca2+-dependent protein kinase C (Jorgensen,
and a small diameter. These fibres exhibit a low excitation thresh- Jensen, & Richter, 2007; Koulmann & Bigard, 2006; Reznick & Shul-
old (high Ca2+ sensitivity) and use a large amount of energy in vivo man, 2006) to modulate numerous regulatory pathways. Calcium
because they are often recruited to sustain low intensity contrac- ions can also modulate glycogen phosphorylase activity (Tate,
tions for basic movements. They are poor in glycogen, rich in myo- Hyek, & Taffet, 1991). Thus, Ca2+ is a highly versatile intracellular
globine and triglycerides, and highly resistant to fatigue. It is well signal that operates over a wide spatiotemporal range to regulate
established that glucose uptake (Bonen, Tan, & Watson-Wright, many different cellular processes in skeletal muscle. Low and sus-
1981), GLUT-4 levels and hexokinase activity (Kern et al., 1990), tained repetitive global Ca2+ transients in slow-twitch fibres and
incorporation of 14C-glucose into glycogen (Bär & Blanchaer, high and rapid localized spikes of Ca2+ in fast-twitch fibres are cen-
1965), turnover of glycogen (Villa-Moruzzi, Locci-Cubeddu, & tral for a selective control of gene expression and metabolic path-
Bergamini, 1979), number of insulin receptors (Lefaucheur, Le ways in different fibre types. For example, only Ca2+ transients
Peuch, Barenton, & Vigneron, 1986), and insulin sensitivity (Henriksen corresponding to those encountered in slow fibres can activate cal-
et al., 1990) are higher in slow-twitch type I than fast-twitch type II cineurin to stimulate the activation of many slow-fibre genes in-
fibres, in particular type IIb glycolytic fibres. On the opposite, gly- volved in contractile process, Ca2+ uptake and energy metabolism
colytic types IIb fibres exhibit high myofibrillar and SR ATPase (Naya et al., 2000). In addition to SR, mitochondria have also been
activities, and can sustain brief and intense contractions fueled shown to actively participate in the regulation of Ca2+ homeostasis
by immediate availability of phosphocreatine and degradation of by rapidly and efficiently accumulating Ca2+ in the mitochondrial
local glycogen through the glycolytic pathway. These fibres exhibit matrix (Brini et al., 1997). Altogether, Ca2+ ions are heteroge-
a high excitation threshold and do not use a large amount of en- neously distributed within skeletal muscle, and Ca2+ flux and/or
ergy in vivo because they are only occasionally recruited to sustain leak is a fundamental process influencing both in vivo muscle biol-
violent movements of short duration, e.g. to escape immediate ogy, as well as post-mortem (p.m.) conversion of muscle to meat.
danger. In association with a low vascularization and a large diam-
eter, glycolytic types IIb fibres are rich in glycogen, poor in myoglo- 2.4.2. Energy state
bine and triglycerides, and highly fatigable. Their low GLUT-4 Because increases in free ADP and AMP levels during contrac-
levels and hexokinase activity denote a predominant utilization tion are higher in fast-twitch glycolytic than slow-twitch oxidative
of locally stored glycogen during intense contraction, with produc- fibres, free ADP and AMP levels are particularly important signals
tion of lactate. Characteristics of type IIx fibres are close to those of for the metabolic control of ATP resynthesis through the anaerobic
IIb fibres, excepted that they speed of contraction is lower and pathway in glycolytic fibres (Hochachka & McClelland, 1997). In-
their oxidative metabolism slightly higher. Type IIa fibres exhibit deed, AMP has been shown to be an allosteric activator of glycogen
intermediate contractile and metabolic properties between types phosphorylase and phosphofructo kinase (Longnus, Wambolt, Par-
I and IIx fibres. sons, Brownsey, & Allard, 2003; Miyamoto et al., 2007). In addition,
Intramyofibre lipids (IMFL), i.e. those contained within myofi- AMP stimulates the activity of AMP-activated protein kinase
bres, can be subdivided into phospholipids, triacylglycerols (AMPK), a fuel gauge monitoring the status of cellular energy
(TAG), mono- and diacylglycerols, cholesterol, cholesteryl esters, (Hardie & Hawley, 2001). AMP binding to the c-subunit of AMPK,
and free fatty acids. The level of TAG (oil red O staining) is high which binding is antagonized by high ATP concentrations, stimulates
in slow-twitch type I fibres, moderate in oxido-glycolytic IIA fibres, AMPK activity both directly by an allosteric mechanism, and indi-
and very low in glycolytic IIB fibres (Essén-Gustavsson, Karlsson, rectly through phosphorylation of AMPK on threonine 172 of the a
Lundstrom, & Enfält, 1994; Malenfant et al., 2001). Phospholipids subunit (Hardie & Sakamoto, 2006; Hawley et al., 1995). Activation
are the main constituents of cellular membranes, and their contri- of AMPK switches on pathways that generate ATP, such as glycol-
bution to total intramuscular fat (IMF) varies between 0.7% and ysis, glucose and fatty acid uptake, fatty acid oxidation, and mito-
0.9%, and is about 30% higher in oxidative than glycolytic muscles, chondrial biogenesis, whereas it switches off ATP-consuming
likely in relation to the higher mitochondrial content of oxidative processes such as glycogen, fatty acid and protein synthesis (Har-
muscles (Alasnier, Rémignon, & Gandemer, 1996; Leseigneur-Mey- die & Hawley, 2001; Jorgensen et al., 2004; Kahn, Alquier, Carling,
nier & Gandemer, 1991). Positive correlations between muscle & Hardie, 2005; Kurth-Kraczek, Hirshman, Goodyear, & Winder,
insulin resistance and accumulation of intramyofibre TAG have 1999; Long & Zierath, 2005; Marsin et al., 2000; Wojtaszewski, Jor-
been reported (van Loon & Goodpaster, 2006). However, this corre- gensen, Hellsten, Hardie, & Richter, 2002). However, the promoting
lation does not seem to represent a functional relationship because effect of AMPK on glucose uptake can induce an accumulation of G-
(1) type I fibres contain more TAG and are more sensitive to insu- 6-P, a potent allosteric activator of glycogen synthase (GS), which
lin, and (2) endurance training increases both intramyofibre TAG seems to override the inhibition of glycogen synthesis induced
L. Lefaucheur / Meat Science 84 (2010) 257–270 261

by phosphorylation of GS by AMPK, thus leading to an accumula- relationships were not always detected (Blanchard, Willis, Warkup,
tion of muscle glycogen. AMPK is a heterotrimeric serine/threonine & Ellis, 2000; Rincker, Killefer, Ellis, Brewer, & Mckeith, 2008; Rinc-
protein kinase composed of a catalytic a-subunit (a1 or a2), and ker, Killefer, Matzat, Carr, & Mckeith, 2009). In pork shops, increas-
regulatory b- (b1 or b2) and c- (c1, c2, and c3) subunits. Expres- ing IMF content up to 2.5–3% is suggested to improve flavor,
sion of a1, b1, c1 and c2 is ubiquitous, a2 and b2 are mostly ex- juiciness and tenderness (Devol et al., 1988), whereas a higher
pressed in skeletal and cardiac muscles and liver, whereas c3 is IMF content has adverse effects on consumer acceptability due to
exclusively expressed in skeletal muscle. Interestingly, c3 is more increased visibility of fat in meat (marbling). Total IMF is com-
highly expressed in fast-twitch glycolytic than oxido-glycolytic fi- posed of triglycerides and phospholipids which represent 0.5–7%
bres as a2b2c3 heterotrimers, whereas it is virtually undetectable and about 0.5% of fresh pig longissimus muscle, respectively
in slow-twitch oxidative fibres (Barnes et al., 2004; Mahlapuu (Faucitano, Rivest, Daigle, Levesque, & Gariepy, 2004; Leseigneur-
et al., 2004; Yu, Fujii, Hirshman, Pomerleau, & Goodyear, 2004). Meynier & Gandemer, 1991). A small amount of triglycerides is
In addition, 5-aminoimidazole-4-carboxamide-1-b-D-ribonucleo- localized as droplets mostly within type I myofibres, whereas an
side (AICAR), a specific AMPK activator, has been shown to have overwhelming amount of triglycerides is located in intramuscular
no effect on AMPK activity, GLUT-4 expression and glucose uptake adipocytes mostly grouped along the perimysium, or interspersed
in the slow-twitch rat soleus muscle (Ai et al., 2002; Balon & Jas- between myofibres (Essén-Gustavsson et al., 1994; Gondret, Mou-
man, 2001; Buhl et al., 2001; Wright, Geiger, Holloszy, & Ho-Han, rot, & Bonneau, 1998). It is often stated that red oxidative muscles
2005). contain more total IMF than white glycolytic muscles. However, a
In addition to LKB1, Ca2+/calmodulin-dependent kinase kinase careful analysis of available data does not confirm this statement,
(CaMKK) would also be an upstream kinase of AMPK (Hawley and shows that the amount and size of intramuscular adipocytes
et al., 2005; Hurley et al., 2005; Woods et al., 2005), suggesting po- within a muscle would not be related to its fibre type composition.
sitive interrelations between Ca2+ and AMPK pathways. However, a This has been shown many times in pig within populations (Eggert,
chronic exposure to elevated cytosolic Ca2+ concentration has re- Depreux, Schinckel, Grant, & Gerrard, 2002; Fiedler, Nurnberg,
cently been shown to block AMPK-induced GLUT-4 expression in Hardge, Nurnberg, & Ender, 2003; Larzul et al., 1997), between
muscle (Park et al., 2009), denoting complex interconnections be- genotypes (Essén-Gustavsson & Fjelkner-Modig, 1985; Lebret
tween Ca2+ and AMPK pathways. et al., 1999) and different muscles (Beecher, Cassens, Hoekstra, &
Briskey, 1965; Leseigneur-Meynier & Gandemer, 1991), and in
feeding experiments (Candek-Potokar, Lefaucheur, Zlender, &
2.4.3. Mitochondrial respiration
Bonneau, 1999; Essén-Gustavsson et al., 1994). Data from our lab-
The numerous mitochondria encountered in oxidative fibres ex-
oratory even show that IMF content can be up to three times higher
hibit a radial gradient distribution with a higher density in the sub-
in the white glycolytic than red oxidative portion of pig semitendi-
sarcolemmal area, whereas the much less abundant mitochondria
nosus muscle, which is consistent with other data reporting that
of glycolytic fibres are homogeneously distributed throughout the
the frequency of IIBW fibres and IMF content can be positively cor-
fibre (Swatland, 1984). Beyond these differences in the amount and
related (Henckel, Oksbjerg, Erlandsen, Barton-Gade, & Bejerholm,
distribution of mitochondria between fibre types, the regulation of
1997). Finally, it is noteworthy that most intramuscular adipocytes
mitochondrial respiration has been shown to differ between fibre
are located in the perimysium, in the vicinity of glycolytic fibres
types. First, respiration is more sensitive to the inhibitory effect
which are the prominent fibre type at the periphery of the myofi-
of physiological concentrations of ATP in mitochondria isolated
bre fascicles. All these data denote no universal relationship be-
from fast- than slow-twitch muscle (Gueguen, Lefaucheur, Ecolan,
tween intramuscular adipocyte content and fibre type
Fillault, & Herpin, 2005). Second, data obtained on permeabilized
composition, and suggest that both characteristics are rather inde-
fibres demonstrate that stimulation of mitochondrial respiration
pendent and can be manipulated separately. On the opposite, a
by ADP is more sensitive in types IIb and IIx (Km = 8 lM) than IIa
strong positive genetic correlation (rg = 0.68) between IMF content
(72 lM) and I (212 lM) fibres, suggesting that mitochondrial res-
and myofibre CSA has been observed in Large White pigs (Larzul
piration is strongly regulated by free ADP in types IIb and IIx fibres,
et al., 1997). In most cases, phenotypic and genetic correlations be-
once ATP level tends to decrease (Gueguen, Lefaucheur, Fillault, &
tween animal general fatness and total IMF content are positive,
Herpin, 2005; Gueguen, Lefaucheur, Fillaut, Vincent, & Herpin,
even though the genetic correlation only reached 0.30 in pig
2005). In contrast, the addition of creatine or AMP strongly stimu-
(Sellier, 1998), showing that IMF is not totally genetically related
lates mitochondrial respiration in type I fibres, and has no effect in
to body fatness (Suzuki et al., 2005).
types IIb and IIx fibres, denoting a strong coupling between mito-
Collagen, one of the main components of muscular connective
chondrial creatine (CK) and adenylate (AK) kinases and oxidative
tissue, plays an important role in binding myofibres together and
phosphorylation in type I fibres, and a weak coupling in glycolytic
into fascicules. In the rat, slow-twitch muscles contain more colla-
fibres, IIA fibres are intermediate. Thus, the high levels of phospho-
gen than fast-twitch muscles (Kovanen, Suominen, & Heikkinen,
creatine and cytosolic MM-CK in glycolytic fibres function as an
1984). However, no clear relationship between collagen content
energy buffer coupled with rapid glycolytic ATP production during
and fibre type composition has been reported in livestock species.
contraction, and mitochondrial ATP regeneration during recovery.
As the animals get older, covalent crosslinks form, linking individ-
On the opposite, the mitochondrial CK, in association with the
ual collagen molecules together and making connective tissue
cytosolic muscle CK (MM-CK) functions as a channelling system
more rigid and less soluble. The increase in collagen concentration
to efficiently transfer ATP and ADP between mitochondria and sites
and the decrease in its solubility are well known to decrease ten-
of ATP consumption via successive phosphocreatine and creatine
derness of meat, but no specific relationship with muscle fibre type
cyclings in type I fibres.
composition has been documented.

2.5. Relationships between myofibre type composition and other


muscle components 3. Significance of myofibre type for growth performance

Intramuscular fat (IMF) content is generally reported to be pos- In most farm mammals and birds, myogenesis is at least a bi-
itively related with meat tenderness, juiciness and flavor (Fernan- phasic phenomenon involving the successive differentiation of a
dez, Monin, Talmant, Mourot, & Lebret, 1999), even though these primary and secondary generation of myotubes during the fetal
262 L. Lefaucheur / Meat Science 84 (2010) 257–270

TNF CSA, µm 2 have been shown to exhibit a lower TNF in semitendinosus mus-
W cle, larger myofibres and a decreased carcass lean meat content at
110 kg BW
8000 commercial slaughter weight than normal birth weight piglets, as
well as increased adiposity and IMF content (Gondret, Lefaucheur,
R
Juin, Louveau, & Lebret, 2006; Hegarty & Allen, 1978; Miller, Gar-
wood, & Judge, 1975; Powell & Aberle, 1981; Rehfeldt & Kuhn,
2006). These differences at equal slaughter weight are likely re-
lated to a higher physiological maturity of low birth weight pig-
100
lets where the plateau of muscle fibre hypertrophy is achieved
90 Age, d 165 d
Age, d earlier because of a lower TNF. Overall, a high TNF combined with
Birth Birth moderate mean myofibre CSA is a way to increase muscle growth
(114 d) potential without decreasing the oxidative capacity which could
Fig. 3. Schematic representation of changes in the total number of fibres (TNF) and
be induced by excessive myofibre hypertrophy. Growth in myofi-
their cross-sectional area (CSA) in future white (W) and red (R) myofibres during bre length is also likely an important mechanism to explain mus-
development in the pig. cle growth and the increase in CSA of some muscles such as that
of longissimus where myofibres make a 25° angle to the vertebral
axis (Davies, 1972). However, growth in length of myofibres has
been poorly studied.
period. The number of secondaries around each primary varies
from between 5 and 9 in mouse and rat (Ontell, Bourke, & Hughes,
1988; Ross, Duxson, & Harris, 1987) to over 20 in larger species, 3.2. Fibre type composition
such as pig (Stickland & Handel, 1986). The total number of fibres
(TNF) is generally considered to be established by 90 dg in the pig, Comparisons between wild and domesticated animals within
i.e. after 80% of gestation (Wigmore & Stickland, 1983), and by different species suggest that selection for increased growth rate
180 dg in cattle, i.e. after 66% of gestation (Picard, Lefaucheur, Ber- and lean meat content has increased myofibre CSA, and shifted
ri, & Duclos, 2002), denoting a higher maturity in cattle. A third muscle metabolism towards a more white glycolytic and less red
generation of fibres forming around some secondaries has been de- oxidative type (Ashmore, 1974; Rahelic & Puac, 1981; Ruusunen
scribed shortly after birth in large species such as sheep, pig, hu- & Puolanne, 2004; Solomon & West, 1985; Weiler, Appell, Kremser,
man and cattle (Ashton, Bayol, Mcentee, Maltby, & Stickland, Hofäcker, & Claus, 1995). A comparison between adult LW
2005; Lefaucheur, Edom, Ecolan, & Butler-Browne, 1995; Masca- (28.2 months, 260 kg) and Meishan pigs (25.6 months, 184 kg)
rello, Stecchini, Rowlerson, & Ballocchi, 1992). Contrary to TNF, supports this hypothesis in domesticated pigs in that LW pigs were
myofibre CSA remains rather constant during gestation, and dra- leaner and exhibited higher TNF, larger IIB glycolytic fibres, and
matically increases postnatally, to a greater extent in future glyco- were more glycolytic and less oxidative than Meishan (Bonneau,
lytic than oxidative fibres (Fig. 3). Contractile and metabolic Mourot, Noblet, Lefaucheur, & Bidanel, 1990). Other comparisons
maturations of myofibres mostly occur either at the end of gesta- between different pig breeds suggests that high muscularity is pos-
tion in cattle (Gagniere, Picard, Jurie, & Geay, 1999), or during itively related with a high abundance of MyHC IIb transcript (Wim-
the first post-natal weeks in less mature species such as pig mers et al., 2008). Within the LW breed, two selection experiments
(Lefaucheur & Vigneron, 1986) and rabbit (Gondret, Lefaucheur, found a positive relationship between the percentage of slow-
Dalbis, & Bonneau, 1996). twitch type I fibres in longissimus muscle and carcass fatness
(Brocks, Hulsegge, & Merkus, 1998; Lefaucheur et al., 2000). A sim-
3.1. Total number of fibres and myofibre cross-sectional area ilar trend was observed between the percentage of oxidative fibres
and carcass fatness in the sheep (Kadim, Purchas, Davies, Rae, &
Muscle mass is related to TNF, myofibre CSA and length of Barton, 1993). However, these results have not been confirmed in
myofibres. Between species variability of TNF is much higher than other experiments carried out in conventional European domesti-
that of fibre CSA (Plaghki, 1985; Rehfeldt, Fiedler, & Stickland, cated pigs (Henckel et al., 1997; Karlsson et al., 1993; Larzul
2004). At a constant body weight, a negative phenotypic correla- et al., 1997; Oksbjerg et al., 2000; Ruusunen, Sevonaimonen, &
tion, between 0.3 to 0.8 in pig, is reported between TNF and Puolanne, 1996; Seideman, Crouse, & Mersmann, 1989; Tribout
CSA (Kim et al., 2008; Rehfeldt, Fiedler, Dietl, & Ender, 2000), et al., 2004). In particular, an increase in carcass lean percentage
and a positive correlation between TNF and lean tissue growth and lean tissue growth rate were both positively related to both
potential is usually reported (Dwyer, Fletcher, & Stickland, 1993; glycolytic and oxidative metabolisms in longissimus (Henckel
Handel & Stickland, 1988; Henckel et al., 1997; Kim et al., 2008; et al., 1997; Karlsson et al., 1993). On the opposite, two experi-
Luff & Goldspink, 1967). The last correlation is very obvious when ments implemented to estimate the realized genetic trends in
comparing double-muscled with normal cattle (Holmes & Ash- French LW pigs from 1977 to 1998 (Tribout et al., 2004), and
more, 1972). Larzul et al. (1997) found a positive genetic correla- Danish Landrace pigs from 1976 to 1995 (Oksbjerg et al., 2000)
tion between lean tissue growth rate and myofibre mean CSA confirmed the expected dramatic improvement in growth rate
(rg = 0.47); however, a careful examination of the literature shows and carcass leanness, but surprisingly found a decreased activity
a controversial relationship between carcass lean meat content of lactate dehydrogenase, a glycolytic enzyme, in longissimus mus-
and fibre CSA. This controversy can be explained by the fact that cle. Finally, other studies did not find any significant correlations
carcass lean content and muscle growth potential would be pri- between lean tissue growth rate, or loin eye area, and fibre type
marily dependent on TNF. Therefore, one should first take into ac- composition in pig longissimus muscle (Candek-Potokar et al.,
count TNF before studying any relationships between muscle 1999; Larzul et al., 1997; Ryu, Rhee, & Kim, 2004). Taken together,
development and myofibre CSA. Indeed, a strong positive correla- strong relationships between fibre type composition and growth
tion between myofibre CSA and muscle development is expected performance can be found when comparing extreme genetic mod-
at a constant TNF. Sometimes, a negative correlation is reported els, such as wild versus domesticated animals, but many of these
between lean meat content and myofibre CSA, likely due to differ- relationships become controversial when studied within conven-
ences in TNF. This is the case in low birth weight piglets which tional domesticated animals.
L. Lefaucheur / Meat Science 84 (2010) 257–270 263

4. Fibre type and meat quality increase in myofibre CSA, sometimes in a fibre type dependant
manner, and influence to variable degrees some meat quality traits,
Following exsanguination, skeletal muscles become isolated such as water holding capacity, colour and sensory quality (Cannon
structures and are no more supplied in oxygen and exogenous et al., 1995). However, this kind of correlations is not sufficient to
nutrients. ATP production is necessary to keep the muscle in the conclude that large myofibre CSA is the causal factor of meat qual-
relaxed state, and the formation of permanent acto-myosin cross- ity deterioration. Indeed, the mechanism leading to meat quality
bridges occur when ATP level decreases, leading to rigor mortis. It deterioration in the mutated Ryr1 pigs is a punctual mutation in
is well documented that p.m. conversion of muscle to meat is ini- the SR Ca2+ channel, leading to an abnormally high release of
tiated by Ca2+ leaking within the muscle and subsequent activation Ca2+ in the sarcoplasm and rapid early p.m. drop in pH, giving rise
of ATPases, modification of adenine nucleotide concentrations, and to Pale Soft and Exudative meat (PSE). In RN carrier pigs, the R225Q
degradation of glycogen to lactic acid through glycolysis. This leads RN single point mutation in the c3-subunit of AMPK increases
to acidification of the meat with variable rate and extent of pH de- AMPK activity (Park et al., 2009), in accordance with data from
crease depending on the rate of H+ production and muscle buffer- similar AMPK mutant mice where the mutation has been shown
ing capacity. In addition to these physicochemical changes, to eliminate the allosteric regulation of AMPK by ATP/AMP, result-
proteolytic systems are stimulated, leading to p.m. maturation of ing in increased basal AMPK activity and muscle glycogen storage
the meat. Meat quality is known to be influenced by numerous fac- (Barnes et al., 2004). This dominant RN mutation is responsible for
tors such as genetic, pre-slaughter and slaughter conditions, p.m. a 70% selective increase of glycogen content in glycolytic muscles,
processing, as well as nutritional and environmental conditions whereas oxidative muscles are unaffected (Estrade, Vignon, Rock, &
during the rearing phase. It is commonly reported that muscle fibre Monin, 1993; Lebret et al., 1999; Marinova, Lefaucheur, Fernandez,
composition influences many aspects of meat quality, including & Monin, 1992; Monin, Mejenes-Quijano, Talmant, & Sellier, 1987),
colour, water holding capacity, tenderness, juiciness and flavor. leading to a low ultimate pH and acid meat with low water holding
However, identifying the specific relationships between meat qual- capacity in glycolytic muscles (Leroy et al., 2000). Mutations and/or
ity and myofibre characteristics, i.e. TNF, myofibre CSA and fibre deletions in the bovine myostatin gene is the causal factor explain-
types, remains a difficult task as discussed in the following ing the double-muscled phenotype mostly due to a dramatic in-
subsections. crease in TNF and, to a lower extend, a specific hypertrophy of
glycolytic fibres (Holmes & Ashmore, 1972; Ouhayoun & Beau-
4.1. Total number of fibres and myofibre cross-sectional area mont, 1968). The superior tenderness of double-muscled cattle
has been attributed to a lower total and higher soluble collagen
As previously reported, TNF and myofibre CSA are frequently content (Ngapo et al., 2002). In Callipyge sheep, a causative point
negatively correlated at a given body weight. In pig, it is widely re- mutation is located in an intergenic region and the mechanism
ported that increasing myofibre CSA would be detrimental for leading to the specific increased CSA of fast myofibre could rather
meat quality, in particular water holding capacity and tenderness be a reduced protein degradation, inducing a lower p.m. proteoly-
(Rehfeldt et al., 2000). Several studies found hypertrophy of fast- sis, maturation and tenderization of meat (Lorenzen et al., 2000). In
twitch oxido-glycolytic fibres to be more specifically detrimental b-agonist treated animals, in particular lambs, the decreased meat
to meat tenderness and water holding capacity (Larzul et al., 1997; tenderness was specifically due to a reduced p.m. maturation be-
Lengerken, Maak, Wicke, Fiedler, & Ender, 1994; Maltin et al., cause of an increased calpastatin activity and decreased proteolysis
1997). Several major genes (Table 2), such as mutated Ryr1, IGF- (Koohmaraie, Shackelford, Muggli-Cockett, & Stone, 1991).
II or RN (‘‘Rendement Napole”) genes in pigs (Essén-Gustavsson, Other studies do not support a negative influence of myofibre
Karlstrom, & Lundstrom, 1992; Lebret et al., 1999; Milan et al., CSA on meat quality. Thus, a strong positive genetic correlation
2000; Van den Maagdenberg, Stinckens, Lefaucheur, Buys, & de (rg = 0.68) was found in longissimus muscle of Large White pigs be-
Smet, 2008), the double-muscling gene in cattle (Grobet et al., tween myofibre CSA and IMF content (Larzul et al., 1997), a char-
1997), the Callipyge gene in sheep (Carpenter, Rice, Cockett, & acteristic generally associated with improved tenderness and
Snowder, 1996; Vuocolo et al., 2007), or treatments with b-ago- flavor of pork. In cattle, the relationship between tenderness and
nists (Dunshea, D’Souza, Pethick, Harper, & Warner, 2005), growth mean fibre CSA is highly controversial (Renand, Picard, Touraille,
hormone (Bonneau, 1991), and steroids (Clancy, Lester, & Roche, Berge, & Lepetit, 2001; Seideman & Crouse, 1986). Surprisingly, a
1986; Fritsche, Solomon, Paroczay, & Rumsey, 2000) all induce an study carried out in broiler breast muscle recently showed a posi-
tive influence of myofibre CSA on meat quality through a decreased
muscle glycogen content, higher ultimate pH, improved water
Table 2
holding capacity and better tenderness (Berri et al., 2007). There-
Effects of some major genes on total number of fibres (TNF), myofibre cross-sectional fore, a universal relationship between myofibre CSA and meat
area (CSA) and percentage glycolytic fibres (% W) of valuable muscles in different quality still remains rather controversial. Because myofibre hyper-
species. =, no effect; nd, not determined. trophy is a balance between protein synthesis and degradation, rel-
Pig Cattle Sheep ative changes in both processes strongly affect myofibre growth
Ryr1 a
RN b
IGF-II c
Myostatin d
Callipygee
rate, as well as meat tenderization/tenderness through modulation
of p.m. proteolysis (Koohmaraie, Kent, Shackelford, Veiseth, &
TNF =, = = =
Wheeler, 2002). At first glance, increasing TNF could be a good
CSA
Red fibres nd = = strategy to simultaneously increase lean meat content while pre-
White fibres = nd serving meat quality by reducing myofibre CSA. However, recent
%W = data in pigs suggest that increasing only TNF would not be suffi-
a
Lindholm, Persson, Jonsson, and Andrén (1977), Essén-Gustavsson et al. (1992), cient, and that there would rather be a range of optimum TNF
Pedersen et al. (2001), Depreux, Grant, and Gerrard (2002). which guarantees both high meat percentage and good meat qual-
b
Lebret et al. (1999). ity at a moderate fibre size (Rehfeldt, Tuchscherer, Hartung, &
c
Van den Maagdenberg et al. (2008). Kuhn, 2007). According to other studies, a high TNF combined with
d
Holmes and Ashmore (1972), Ashmore, Parker, Strokes, and Doerr (1974),
Swatland and Kieffer (1974), West (1974).
a low percentage of type IIBW fibres would be a good strategy to
e
Koohmaraie, Shackelford, Wheeler, Lonergan, and Doumit (1995), Carpenter improve both lean meat content and meat quality in pig (Kim
et al. (1996). et al., 2008; Ryu, Lee, Lee, & Kim, 2006).
264 L. Lefaucheur / Meat Science 84 (2010) 257–270

4.2. Fibre type composition confirmed in other normal cattle breeds (Wegner et al., 2000) or
lambs (Solomon & Lynch, 1988).
Because fibre type composition and meat quality can be dra- Increasing the proportion of type I fibres has been reported to
matically different between muscles within the body, a possible decrease the rate and extent of p.m. pH decline and lightness,
association between fibre type characteristics and meat quality and improve water holding capacity in pig (Choi et al., 2006; Gil
has been postulated. However, identifying the optimum fibre type et al., 2003; Larzul et al., 1997; Ryu & Kim, 2005). Increasing the
composition for best technological and sensory quality of meat re- proportion of slow-twitch type I fibres has been reported to im-
mains a challenge. Both the contractile and metabolic nature of fi- prove tenderness and juiciness in cattle (Maltin et al., 1998), and
bres likely influence meat quality. Thus, as previously presented in juiciness and flavor in sheep (Valin, Touraille, Vigneron, & Ash-
the document (Table 1), fast-twitch glycolytic fibres exhibit a more more, 1982). The positive relationship between the proportion of
extensive and efficient SR, higher acto-myosin-ATPase activity, and type I fibres and flavor is likely due to their high level of phospho-
higher levels of glycogen than slow-twitch oxidative fibres which, lipids that have been shown to be major determinants of cooked
conversely, contain a higher number of mitochondria, as well as meat flavor (Meynier & Gandemer, 1994). However, the high levels
more phospholipids and myoglobin. Consequently, according to of polyunsaturated fatty acids in phospholipids also increase the
their composition in fibre types, muscles have different abilities risks of p.m. fatty acid oxidation and development of rancidity.
to release and sequester Ca2+, activate ATPase activities, produce Moreover, a high proportion of type I fibres is prone to the produc-
adenine nucleotides, stimulate glycolysis, produce lactate and de- tion of dark, firm, and dry meat (DFD), in particular in beef meat
crease p.m. pH depending on their buffering capacity (Bowker, (Ozawa et al., 1999). It is also well documented that increasing
Grant, Swartz, & Gerrard, 2004). Consistent with this, increasing the proportion of type I fibres also decreases colour stability with
the proportion of fast-twitch glycolytic fibres (IIBW) in longissimus a possible shift to a brownish stain (Renerre, 1984).
muscle has been shown to increase the rate and extent of p.m. pH Taken together, some relationships between fibre type compo-
decline, lightness (L*), paleness, cooking loss, protein denaturation, sition and meat quality do exist but are not universal among spe-
and decrease water holding capacity in pigs (Choe et al., 2008; cies and depend on numerous interacting factors such as muscle
Choi, Ryu, & Kim, 2006; Choi, Ryu, & Kim, 2007; Kauffman et al., type, species, breed, genotype, age, nutrition, environmental
1998; Larzul et al., 1997; Rosenvold et al., 2001; Ryu & Kim, factors, as well as slaughter conditions and p.m. processing. In
2005; Ryu & Kim, 2006; Ryu et al., 2008). However, other less addition to myofibres, other muscle components such as intramus-
numerous studies have also reported that increasing the propor- cular adipocytes and connective tissue can influence multiple
tion of fast-twitch oxido-glycolytic fibres decreased water holding aspects of meat quality in an interconnected manner. Because of
capacity of cooked meat, tenderness, juiciness and flavor (Henckel this complex network of interactions, it is difficult to find a simple
et al., 1997; Maltin et al., 1997). Interestingly, a comparison be- and universal biological marker of meat quality that would be
tween different pig muscles showed that the rate of p.m. pH de- usable in all situations. Rather, a better control of meat quality
cline was much faster in the mixed psoas major than both the needs a good understanding of muscle biology as a global complex
slow-twitch oxidative semispinalis and fast-twitch glycolytic lon- system submitted to the influence of multiple intrinsic and extrin-
gissimus (Fig. 4), showing that rate of p.m. pH decline can depend sic factors.
on other factors than fibre type composition, such as muscle buf-
fering capacity.
5. Tools to control meat quality through fibre type composition
Post-mortem rate of maturation has been shown to be influ-
enced by fibre type composition. Thus, a study including different
Muscle fibres are dynamic structures which exhibit high plas-
muscles in cattle, pigs and sheep reported a higher calpain/calpast-
ticity and undergo type shift following an obligatory pathway
atin ratio in fast-twitch glycolytic than slow-twitch oxidative mus-
I M IIa M IIx M IIb (Pette & Staron, 2000; Schiaffino & Reggiani,
cles (Ouali & Talmant, 1990), which is consistent with the faster
1994). Muscle histochemical characteristics can be influenced by
p.m. rate of maturation in fast than slow muscles (Jurie, Robelin,
numerous intrinsic and extrinsic factors such as muscle location,
Picard, & Geay, 1995; Ouali, 1990). Therefore, increasing the pro-
species, breed, genotype, gender, age, exercise, ambient tempera-
portion of fast-twitch glycolytic fibres could have beneficial effects
ture, nutrition, and growth promoting agents (Gondret, Combes,
on p.m. maturation and tenderness in species exhibiting slow p.m.
Lefaucheur, & Lebret, 2005; Gunning & Hardeman, 1991; Pette &
maturation, such as cattle (Seideman, Crouse, & Cross, 1986;
Staron, 2001; Swynghedauw, 1986). Meat quality is also influenced
Zamora et al., 1996). This is in agreement with others results in cat-
by many factors, among which the most important ones are re-
tle showing a better tenderness in the superficial more glycolytic
ported to be muscle location, species, pre-slaughter and slaughter
than deep more oxidative region of semitendinosus muscle (Tot-
conditions, post-slaughter processing, genetic factors, growth
land, Kryvi, & Slinde, 1988). However, such correlations were not
promoting agents, age and weight (Andersen, Oksbjerg, Young, &
Therkildsen, 2005; Dunshea, D’Souza, Pethick, Harper, & Warner,
2005; Ngapo & Gariepy, 2008; Rosenvold & Andersen, 2003;
I Semispinalis (S) Scheffler & Gerrard, 2007). Despite significant effects of some
pH 24h
100 growth promoting agents on both myofibre characteristics and
IIA Psoas major (P)
80 6.5 Longissimus (L) meat quality, this topic will not be presented in the present review
IIB
because the choice was made to focus on some factors susceptible
60
% to be manipulated in conventional rearing conditions, i.e. genetic
40 6 factors, physical exercise, ambient temperature and nutrition.
20
0 5.5 5.1. Genetic factors
S P L 45 min p.m. 24h p.m.
Within species, muscle fibre type composition is strongly influ-
Fig. 4. Fibre type composition using conventional histochemical mATPase staining
enced by genetic factors, such as breed, genotype and presence of
after preincubation at pH 4.35 (left) and post-mortem pH (right) in semispinalis (S),
psoas major (P) and longissimus (L) muscles in the Large White pig. Lefaucheur major genes. However, because differences between breeds are
et al. (unpublished). also related to other factors than fibre type composition, i.e.
L. Lefaucheur / Meat Science 84 (2010) 257–270 265

growth rate, body composition, and IMF content, these models are 50
not well suited to specifically address the relationships between
myofibre characteristics, growth performance, and meat quality. 40

Fiber number, %
This limitation is obvious for major genes presented in Table 2. In-
deed, they all induce an increase in carcass lean content, with var- 30 *
iable effects on TNF, myofibre CSA and fibre type composition not
directly related to differences in meat quality. Only the mutated 20
myostatin gene in cattle is associated with an additional increase **
in TNF. On the opposite, increased muscle content of Ryr1 pigs is 10
exclusively related to increased myofibre CSA because they have
been shown to have either similar or reduced TNF. Morphologi-
0
cally, RN carrier pigs only exhibit an increased myofibre CSA of
I IIA IIBR IIBW
fast-twitch oxido-glycolytic fibres. Interestingly, RN carrier pigs
also show increased glycogen level, AMPK and acetyl-CoA carbox- Fig. 6. Fibre type composition of triceps brachii muscle from sedentary (white bars)
ylase phosphorylations, AMPK activity, GLUT-4 expression at the and trained (black bars) miniature pigs. Animals were trained 5 d/week over 20
mRNA and protein levels, PGC-1a and FAT/CD36 mRNA levels in week. *, P < 0.05; **, P < 0.01). Adapted from McAllister et al. (1997).
longissimus muscle compared to wild type pigs (Park et al.,
2009), suggesting a greater reliance of glycolytic muscles on fatty to be a promising target to influence both myofibre characteristics
acid metabolism in mutated RN pigs. This is in accordance with and meat quality.
previous data showing an enhanced oxidative metabolism, CSA Fibre type composition is highly variable between individuals of
and relative area of type IIA and IIBR oxydo-glycolytic fibres, and the same breed reared under similar nutritional and environmental
a decreased glycolytic metabolism in longissimus muscle of RN conditions. Within the LW pig breed, the percentage of type I fibres
carrier pigs (Lebret et al., 1999). Even though total IMF content in the longissimus muscle ranges from 1.6% to 21.5% (9.5 ± 2.66,
of longissimus was not influenced by the RN genotype in this last CV = 28%), and mean myofibre CSA from 2560 to 6040 lm2
study, IMFL could have been decreased as shown in vastus lateralis (3410 ± 620, CV = 18%) (Larzul et al., 1997). The TNF has also been
muscle of humans carrying the AMPK-c3 R225W mutation, which shown to be variable, ranging from 292  103 to 958  103
also lead to a dramatic increase in muscle glycogen content (Cost- (533  103 ± 94  103, CV = 18%) in the semitendinosus muscle
ford et al., 2007). Taken together, most available data support a within an F2 population of 364 LW  Meishan pigs (Lefaucheur
negative correlation between IMFL and glycogen contents at the et al., unpublished). Muscle fibre traits exhibit moderate heritabil-
myofibre level. Interestingly, additional mutated alleles in the ities with h2 values ranging from 0.2 to 0.5 for TNF, 0.2 to 0.3 for
AMPKc3 gene with different effects on muscle glycogen content myofibre mean CSA, around 0.4 for fibre type frequencies, and
have been discovered. Thus, the mutated V199I allele is correlated 0.3 to 0.4 for glycogen in pig longissimus muscle (Larzul et al.,
with lower muscle glycogen and lactate, higher ham and loin pH, 1997, 1999; Oksbjerg, Henckel, Andersen, Pedersen, & Nielsen,
and improved meat colour score, resulting in better meat quality 2004; Rehfeldt et al., 2000). As a comparison, average heritabilities
in pig (Ciobanu et al., 2001). From a practical point of view, manip- are 0.5 for carcass lean percentage, 0.5 for muscle IMF content, and
ulations of these major genes are potential tools to control myofi- 0.1–0.3 for meat quality traits, with 0.25–0.30 for colour and ten-
bre development and meat quality in different species. In derness, and 0.15–0.20 for pH and water holding capacity (Sellier,
particular, because of its central role (1) in vivo in the regulation 1998). The high individual variability of myofibre traits associated
of systemic energy balance (food intake, glucose and lipid homeo- to their moderate heritability makes it possible to select animals
stasis) and skeletal muscle metabolism (glycogen and IMFL con- directly on myofibre traits. This has been done for myofibre CSA
tents, energy metabolism), and (2) p.m. in its involvement in the in pigs (Wicke, Von Lengerken, Fiedler, Altmann, & Ender, 1991);
conversion of muscle to meat (Scheffler & Gerrard, 2007; Shen, however, results were flawed because of the indirect selection of
Gerrard, & Du, 2008; Underwood et al., 2007, 2008), AMPK appears the mutated Ryr1 gene positively correlated with myofibre CSA,
but known to alter meat quality through another mechanism, i.e.
Ca2+ flux disregulation. A direct selection carried out for six gener-
ations against in vivo glycogen levels in longissimus muscle of LW
pigs free of the mutated Ryr1 and RN genes has been shown to in-
duce a decrease in carcass lean content and CSA of fast-twitch oxi-
do-glycolytic fibres, and improve meat technological quality
without any change in IMF content (Larzul et al., 1999). It is note-
1.6
worthy that correlated changes with this selection against muscle
1.4
glycogen content were consistent with the effects of the RN gene
[gly co ge n], %

1.2 on carcass lean content, muscle glycogen level, myofibre character-


1.0 istics and meat quality. The high heritability reported for the per-
0.8 centage of type I fibres in longissimus muscle (h2 = 0.46 ± 0.11)
0.6 (Larzul et al., 1997) led us to start a divergent selection on this
0.4 myofibre trait within the same population of LW pigs. Preliminary
0.2
*** results showed that divergent selection was effective from the first
0.0
* IIBRIIBW generation (h2 = 0.29, 1.3 genetic standard deviation between the
Low RFI IIA two lines) (Lefaucheur et al., 2000). Unfortunately, this experiment
High RFI I had to be stopped after the first generation of selection for sanitary
reasons unrelated to the selection experiment. At present time, a
divergent selection on residual feed intake (RFI) is under way
Fig. 5. Glycogen content of myofibres at 108 kg BW in longissimus muscle biopsies
within the LW breed. After four generations, the animals from
of Large White pigs divergently selected for low and high residual feed intake (RFI).
Myofibres were classified as types I, IIA, IIBR and IIBW. *, P < 0.05; ***, P < 0.001. the RFI-line, i.e. efficient animals, exhibited leaner carcasses and
Adapted from Lefaucheur et al. (2008). lower longissimus IMF content (Lefaucheur et al., 2008). Moreover,
266 L. Lefaucheur / Meat Science 84 (2010) 257–270

RFI-pigs had a higher percentage of IIBW fibres, larger fibres, and when IIb MyHC is expressed. Because measuring histological traits
increased glycogen content, in accordance with a lower ultimate is laborious and time consuming, identification of biochemical
pH and reduced water holding capacity. Interestingly, muscle gly- markers of these traits would be a breakthrough. Unfortunately,
cogen content increased specifically in type IIBW fibres (Fig. 5) and the identification of such markers remains elusive. Thus, poor cor-
further work is needed to identify the underlying mechanisms. relations between serum endocrine factor levels and muscle fibre
type composition have been found (Ryu, Choi, Ko, & Kim, 2007).
5.2. Exercise, ambient temperature and nutrition In the future, integrative techniques such as genomic, transcrip-
tomic, proteomic, and metabonomic could help to find such mark-
Physical exercise is well known to influence myofibre charac- ers. So far, most significant advances have been done through the
teristics, depending on the type and duration of the exercise. Thus, identification of QTL but very few of them relate to myofibre char-
endurance exercise induces a IIBW ? IIBR ? IIA ? I transition in acteristics (Estelle et al., 2008; Nii et al., 2005; Wimmers, Murani,
muscles involved in the exercise. This has also been found in min- Ngu, Schellander, & Ponsuksili, 2007), suggesting that myofibre
iature pigs (Mcallister, Reiter, Amann, & Laughlin, 1997) (Fig. 6), traits are under the control of multiple genes, in interaction with
and confirmed in more conventional pigs, even though changes environmental factors.
were much weaker and probably not always of physiological Numerous studies identified TNF as an important factor in-
importance (Essén-Gustavsson, 1993; Petersen, Henckel, Oksbjerg, volved in the control of both muscle growth and meat quality,
& Sorensen, 1998). Long term cold exposure has been shown to but its measurement remains a difficult task. Because TNF is fixed
shift fibre type composition to a slower type in oxidative muscles before birth in most livestock animals, further work will have to be
involved in posture in LW pigs (Herpin & Lefaucheur, 1992; Lefauc- done to study the influence of prenatal events on myogenesis. Due
heur et al., 1991). At similar growth rates, an exposure to 12 °C vs. to the high genetic variability and rather good heritability of most
28 °C between 8 and 92 kg BW also increased the rate and extent of myofibre traits, selection experiments directly based on myofibre
p.m. pH decline in longissimus, and enhanced the problem of bico- characteristics within breeds, and the study of correlated effects
lourism between muscles, which likely alter meat quality. On the on growth performance and meat quality parameters are promis-
opposite, exposure to a warm environment has been reported to ing ways to more specifically identify the relationships between
decrease both oxidative and glycolytic metabolisms in pig glyco- myofibre characteristics and meat quality. In addition to the
lytic muscles (Rinaldo & Le Dividich, 1991). Finally, 50% feed in vivo approaches, in vitro studies based on muscle cell cultures
restriction at thermoneutrality between 3 and 7 week of age did should complete the experimental design to better understand
not change fibre type proportions in longissimus muscle, whereas the mechanisms involved in the variability of TNF, myofibre CSA
it induced a 43% increase in the percentage of type I fibres in rhom- and fibre type composition in meat producing animals. These
boideus muscle, a red neck muscle (Harrison, Rowlerson, & Daun- in vitro systems will be invaluable tools to study the regulation
cey, 1996). Taken together, physical exercise, ambient temperature of genes involved in the determination of fibre type characteristics,
and nutrition are potential tools which can be combined with ge- such as those coding for myostatin and the different MyHC or those
netic factors (breed, genotype) in different rearing systems to con- involved in the AMPK network. The development and use of trans-
trol myofibre characteristics and meat quality. genic and knocked-out animals should also help to go further in
the understanding of muscle biology applied to meat production.
6. Conclusion and future perspectives Of course, the achievement of these goals will need a multidisci-
plinary approach combining the competences of meat and animal
The present review shows that some relationships do exist be- scientists, geneticists, biologists, physiologists, statisticians, and
tween myofibre characteristics and meat quality traits under given mathematicians.
conditions. Thus, in pig muscle, increasing the proportion and CSA
of IIBW fibres has often been shown to decrease water holding References
capacity and tenderness. On the opposite, increasing the propor-
Acevedo, L. M., & Rivero, J. L. (2006). New insights into skeletal muscle fibre types in
tion of type IIBW fibres has also been reported to improve meat the dog with particular focus towards hybrid myosin phenotypes. Cell and Tissue
tenderness through improved p.m. maturation in cattle. As evi- Research, 323, 283–303.
denced by the present review, there seems to be no universal rela- Ai, H., Ihlemann, J., Hellsten, Y., Lauritzen, H. P. M. M., Hardie, D. G., Galbo, H., et al.
(2002). Effect of fiber type and nutritional state on AICAR- and contraction-
tionship between myofibre characteristics and meat quality traits stimulated glucose transport in rat muscle. American Journal of Physiology –
among species because both traits are influenced by multiple fac- Endocrinology and Metabolism, 282, E1291–E1300.
tors such as species, breeds, genders, muscle location, sampling Alasnier, C., Rémignon, H., & Gandemer, G. (1996). Lipid characteristics associated
with oxidative and glycolytic fibres in rabbit muscles. Meat Science, 43,
sites within muscle, amount and nature of connective tissue and
213–224.
intramuscular adipocytes, age, slaughter weight, degree of physio- Andersen, H. J., Oksbjerg, N., Young, J. F., & Therkildsen, M. (2005). Feeding and meat
logical maturity at the time of slaughter, birth weight, myofibre quality – A future approach. Meat Science, 70, 543–554.
typing methods and methodology used to measure meat quality. Arguello, A., LopezFernandez, J. L., & Rivero, J. L. L. (2001). Limb myosin heavy chain
isoproteins and muscle fiber types in the adult goat (Capra hircus). Anatomical
Thus, only a global approach including all parameters seems to Record, 264, 284–293.
be pertinent to understand relationships between myofibre char- Ashmore, C. R. (1974). Phenotypic expression of muscle fiber types and some
acteristics, growth performance and meat quality. implications to meat quality. Journal of Animal Science, 38, 1158–1164.
Ashmore, C. R., & Doerr, L. (1971). Comparative aspects of muscle fiber types in
The presence of four adult MyHC, i.e. I, IIa, IIx and IIb shows that different species. Experimental Neurology, 31, 408–418.
conventional enzyme histochemical classifications in three types, Ashmore, C. R., Parker, W., Strokes, H., & Doerr, L. (1974). Comparative aspects of
i.e. I, IIA and IIB must be revisited, at least in species where IIb muscle fibre types in fetuses of the normal and ‘‘double-muscled” cattle.
Growth, 38, 501–506.
MyHC is expressed such as pig, llama and rabbit. Fibre typing could Ashton, C., Bayol, S., Mcentee, G., Maltby, V., & Stickland, N. (2005). Prenatal
be ideally performed using specific monoclonal antibodies raised influences on skeletal muscle development in mammals, birds and fish. Archiv
against each MyHC. Unfortunately, such antibodies are not all fur Tierzucht, 18, 4–10 (Special Issue).
Bacou, F., & Vigneron, P. (1988). Propriétés des fibres musculaires squelettiques. 1.
available and usable with the different immunological techniques Influence de l’innervation motrice. Reproduction Nutrition Dévelopment, 28, 1387.
(immunocytochemistry, western blot, ELISA) in the different spe- Balon, T. W., & Jasman, A. P. (2001). Acute exposure to AICAR increases glucose
cies at different ages. The physical separation of all MyHC by elec- transport in mouse EDL and soleus muscle. Biochemical and Biophysical Research
Communications, 282, 1008–1011.
trophoresis still remains a difficult task in meat producing animals
L. Lefaucheur / Meat Science 84 (2010) 257–270 267

Bär, U., & Blanchaer, M. C. (1965). Glycogen and CO2 production from glucose and Davoli, R., Fontanesi, L., Zambonelli, P., Bigi, D., Gellin, J., Yerle, M., et al. (2002).
lactate by red and white skeletal muscle. American Journal of Physiology, 209, Isolation of porcine expressed sequence tags for the construction of a first
905–909. genomic transcript map of the skeletal muscle in pig. Animal Genetics, 33,
Bär, A., & Pette, D. (1988). Three fast myosin heavy chains in adult rat skeletal 3–18.
muscle. FEBS Letters, 235, 153–155. Depreux, F. F. S., Grant, A. L., & Gerrard, D. E. (2002). Influence of halothane genotype
Barnes, B. R., Marklund, S., Steiler, T. L., Walter, M., Hjalm, G., Amarger, V., et al. and body-weight on myosin heavy chain composition in pig muscle as related
(2004). The 50 -AMP-activated protein kinase gamma 3 isoform has a key role in to meat quality. Livestock Production Science, 73, 265–273.
carbohydrate and lipid metabolism in glycolytic skeletal muscle. Journal of Devol, D. L., Mckeith, F. K., Bechtel, P. J., Novakofski, P. J., Shanks, R. D., & Carr, T. R.
Biological Chemistry, 279, 38441–38447. (1988). Variation in composition and palatability traits and relationships
Beecher, G. R., Cassens, R. G., Hoekstra, W. G., & Briskey, E. J. (1965). Red and white between muscle characteristics and palatability in a random sample of pork
fiber content and associated post-mortem properties of seven porcine muscles. carcasses. Journal of Animal Science, 66, 385–395.
Journal of Food Science, 30, 969–976. Dunshea, F. R., D’Souza, D. N., Pethick, D. W., Harper, G. S., & Warner, R. D. (2005).
Berri, C., Bihan-Duval, E., Debut, M., Sante-Lhoutellier, V., Baeza, E., Gigaud, V., et al. Effects of dietary factors and other metabolic modifiers on quality and
(2007). Consequence of muscle hypertrophy on characteristics of Pectoralis nutritional value of meat. Meat Science, 71, 8–38.
major muscle and breast meat quality of broiler chickens. Journal of Animal Dwyer, C. M., Fletcher, J. M., & Stickland, N. C. (1993). Muscle cellularity and
Science, 85, 2005–2011. postnatal growth in the pig. Journal of Animal Science, 71, 3339–3343.
Blanchard, P. J., Willis, M. B., Warkup, C. C., & Ellis, M. (2000). The influence of Eggert, J. M., Depreux, F. F. S., Schinckel, A. P., Grant, A. L., & Gerrard, D. E. (2002).
carcass backfat and intramuscular fat level on pork eating quality. Journal of Myosin heavy chain isoforms account for variation in pork quality. Meat Science,
Science and Food Agriculture, 80, 145–151. 61, 117–126.
Bonen, A., Tan, M. N., & Watson-Wright, W. M. (1981). Insulin binding and glucose Ennion, S., Sant’Ana Pereira, J., Sargeant, A. J., Young, A., & Goldspink, G. (1995).
uptake differences in rodent skeletal muscles. Diabetes, 30, 702–704. Characterization of human skeletal muscle fibres according to the myosin heavy
Bonneau, M. (1991). Regulation of pig growth by somatotropic hormones: II. The chains they express. Journal of Muscle Research and Cell Motility, 16, 35–43.
effect of exogenous GRF or pST administration on performance and meat Essén-Gustavsson, B. (1993). Muscle-fibre characteristics in pigs and relationships
quality. Pig News Information, 12, 39–45. to meat-quality parameters – Review. In E. Puolanne, D. I. Demeyer, M.
Bonneau, M., Mourot, J., Noblet, J., Lefaucheur, L., & Bidanel, J. P. (1990). Tissue Ruusunen, & S. Ellis (Eds.), Pork quality: Genetic and metabolic factors
development in Meishan pigs: Muscle and fat development and metabolism (pp. 140–159). Wallingford, UK: CAB International.
and growth regulation by somatotropic hormone. In M. Molenat & C. Legault Essén-Gustavsson, B., & Fjelkner-Modig, S. (1985). Skeletal muscle characteristics in
(Eds.), Chinese pig symposium (pp. 203–213). Jouy-en-Josas, France: INRA Press. different breeds of pigs in relation to sensory properties of meat. Meat Science,
Bowker, B. C., Grant, A., Swartz, D. R., & Gerrard, D. E. (2004). Myosin heavy chain 13, 33–47.
isoforms influence myofibrillar ATPase activity under simulated postmortem Essén-Gustavsson, B., Karlsson, A., Lundstrom, K., & Enfält, A. C. (1994).
pH, calcium, and temperature conditions. Meat Science, 67, 139–147. Intramuscular fat and muscle fibre lipid contents in halothane-gene-free pigs
Brini, M., De Giorgi, F., Murgia, M., Marsault, R., Massimino, M. L., Cantini, M., et al. fed high or low protein diets and its relation to meat quality. Meat Science, 38,
(1997). Subcellular analysis of Ca2+ homeostasis in primary cultures of skeletal 269–277.
muscle myotubes. Molecular Biology of the Cell, 8, 129–143. Essén-Gustavsson, B., Karlstrom, K., & Lundstrom, K. (1992). Muscle fibre
Brocks, L., Hulsegge, B., & Merkus, G. (1998). Histochemical characteristics in characteristics and metabolic response at slaughter in pigs of different
relation to meat quality properties in the longissimus lumborum of fast and halothane genotypes and their relation to meat quality. Meat Science, 31, 1–11.
lean growing lines of large white pigs. Meat Science, 50, 411–420. Estelle, J., Gil, F., Vazquez, J. M., Latorre, R., Ramirez, G., Barragan, M. C., et al. (2008).
Brooke, M. H., & Kaiser, K. K. (1970). Muscle fiber types: How many and what kind? A quantitative trait locus genome scan for porcine muscle fiber traits reveals
Archives of Neurology, 23, 369–379. overdominance and epistasis. Journal of Animal Science, 86, 3290–3299.
Buhl, E. S., Jessen, N., Schmitz, O., Pedersen, S. B., Pedersen, O., Holman, G. D., et al. Estrade, M., Vignon, X., Rock, E., & Monin, G. (1993). Glycogen hyperaccumulation in
(2001). Chronic treatment with 5-aminoimidazole-4-carboxamide-1-beta-D- white muscle fibres of RN carrier pigs. A biochemical and ultrastructural study.
ribofuranoside increases insulin-stimulated glucose uptake and GLUT4 Comparative Biochemistry and Physiology Part B: Biochemistry and Molecular
translocation in rat skeletal muscles in a fiber type-specific manner. Diabetes, Biology, 104B, 321–326.
50, 12–17. Faucitano, L., Rivest, J., Daigle, J. P., Levesque, J., & Gariepy, C. (2004). Distribution of
Cameron, N. D., Nute, G. R., Brown, S. N., Enser, M., & Wood, J. D. (1999). Meat intramuscular fat content and marbling within the longissimus muscle of pigs.
quality of Large White pig genotypes selected for components of efficient lean Canadian Journal of Animal Science, 84, 57–61.
growth rate. Animal Science, 68, 115–127. Fernandez, X., Monin, G., Talmant, A., Mourot, J., & Lebret, B. (1999). Influence of
Candek-Potokar, M., Lefaucheur, L., Zlender, B., & Bonneau, M. (1999). Effect of intramuscular fat content on the quality of pig meat – 1. Composition of the
slaughter weight and/or age on histological characteristics of pig longissimus lipid fraction and sensory characteristics of m. longissimus lumborum. Meat
dorsi muscle as related to meat quality. Meat Science, 52, 195–203. Science, 53, 59–65.
Cannon, J. E., Morgan, J. B., Heavner, J., Mckeith, F. K., Smith, G. C., & Meeker, D. L. Fiedler, I., Nurnberg, K., Hardge, T., Nurnberg, G., & Ender, K. (2003). Phenotypic
(1995). Pork quality audit: A review of the factors influencing pork quality. variations of muscle fibre and intramuscular fat traits in Longissimus muscle of
Journal of Muscle Foods, 6, 369–402. F-2 population Duroc  Berlin Miniature Pig and relationships to meat quality.
Carpenter, C. E., Rice, O. D., Cockett, N. E., & Snowder, G. D. (1996). Histology and Meat Science, 63, 131–139.
composition of muscles from normal and Callipyge lambs. Journal of Animal Fritsche, S., Solomon, M. B., Paroczay, E. W., & Rumsey, T. S. (2000). Effects of
Science, 74, 388–393. growth-promoting implants on morphology of longissimus and semitendinosus
Choe, J. H., Choi, Y. M., Lee, S. H., Shin, H. G., Ryu, Y. C., Hong, K. C., et al. (2008). The muscles in finishing steers. Meat Science, 56, 229–237.
relation between glycogen, lactate content and muscle fiber type composition, Gagniere, H., Picard, B., Jurie, C., & Geay, Y. (1999). Comparison of foetal metabolic
and their influence on postmortem glycolytic rate and pork quality. Meat differentiation in three cattle muscles. Reproduction Nutrition Development, 39,
Science, 80, 355–362. 105–112.
Choi, Y. M., Ryu, Y. C., & Kim, B. C. (2006). Effect of myosin heavy chain isoforms on Gil, M., Oliver, M. A., Gispert, M., Diestre, A., Sosnicki, A. A., Lacoste, A., et al. (2003).
muscle fiber characteristics and meat quality in porcine longissimus muscle. The relationship between pig genetics, myosin heavy chain I, biochemical traits
Journal of Muscle Foods, 17, 413–427. and quality of M. Longissimus thoracis. Meat Science, 65, 1063–1070.
Choi, Y. M., Ryu, Y. C., & Kim, B. C. (2007). Influence of myosin heavy- and light chain Gondret, F., Combes, S., Lefaucheur, L., & Lebret, B. (2005). Effects of exercise during
isoforms on early postmortem glycolytic rate and pork quality. Meat Science, 76, growth and alternative rearing systems on muscle fibers and collagen
281–288. properties. Reproduction Nutrition Development, 45, 69–86.
Ciobanu, D., Bastiaansen, J., Malek, M., Helm, J., Woollard, J., Plastow, G., et al. Gondret, F., Lefaucheur, L., Dalbis, A., & Bonneau, M. (1996). Myosin isoform
(2001). Evidence for new alleles in the protein kinase adenosine transitions in four rabbit muscles during postnatal growth. Journal of Muscle
monophosphate-activated gamma(3)-subunit gene associated with low Research and Cell Motility, 17, 657–667.
glycogen content in pig skeletal muscle and improved meat quality. Genetics, Gondret, F., Lefaucheur, L., Juin, H., Louveau, I., & Lebret, B. (2006). Low birth weight
159, 1151–1162. is associated with enlarged muscle fiber area and impaired meat tenderness of
Clancy, M. J., Lester, J. M., & Roche, J. F. (1986). The effects of anabolic agents and the longissimus muscle in pigs. Journal of Animal Science, 84, 93–103.
breed on the fibers of the longissimus muscle of male cattle. Journal of Animal Gondret, F., Mourot, J., & Bonneau, M. (1998). Comparison of intramuscular adipose
Science, 63, 83–91. tissue cellularity in muscles differing in their lipid content and fibre type
Costford, S. R., Kavaslar, N., Ahituv, N., Chaudhry, S. N., Schackwitz, W. S., Dent, R., composition during rabbit growth. Livestock Production Science, 54, 1–10.
et al. (2007). Gain-of-function R225W mutation in human AMPKgamma3 Graziotti, G. H., Palencia, P., Delhon, G., & Rivero, J. L. L. (2004). Neuromuscular
causing increased glycogen and decreased triglyceride in skeletal muscle. PLoS partitioning, architectural design, and myosin fiber types of the M. Vastus
ONE, 2, e903. lateralis of the llama (Lama glama). Journal of Morphology, 262, 667–681.
Cox, R. D., & Buckingham, M. E. (1992). Actin and myosin genes are transcriptionally Graziotti, G. H., Rios, C. M., & Rivero, J. L. L. (2001). Evidence for three fast myosin
regulated during mouse skeletal muscle development. Developmental Biology, heavy chain isoforms in type II skeletal muscle fibers in the adult llama (Lama
149, 228–234. glama). Journal of Histochemistry and Cytochemistry, 49, 1033–1044.
Da Costa, N., Blackley, R., Alzuherri, H., & Chang, K. C. (2002). Quantifying the Grobet, L., Martin, L. J. R., Poncelet, D., Pirottin, D., Brouwers, B., Riquet, J., et al.
temporospatial expression of postnatal porcine skeletal myosin heavy chain (1997). A deletion in the bovine myostatin gene causes the double-muscled
genes. Journal of Histochemistry and Cytochemistry, 50, 353–364. phenotype in cattle. Nature Genetics, 17, 71–74.
Davies, A. S. (1972). Postnatal changes in the histochemical fibre types of porcine Gueguen, N., Lefaucheur, L., Ecolan, P., Fillault, M., & Herpin, P. (2005). Ca2+-
skeletal muscle. Journal of Anatomy, 113, 213–240. activated myosin-ATPases, creatine and adenylate kinases regulate
268 L. Lefaucheur / Meat Science 84 (2010) 257–270

mitochondrial function according to myofibre type in rabbit. Journal of Kern, M., Wells, J. A., Stephens, J. M., Elton, C. W., Friedman, J. E., Tapscott, E. B., et al.
Physiology, 564, 723–735. (1990). Insulin responsiveness in skeletal muscle is determined by glucose
Gueguen, N., Lefaucheur, L., Fillault, M., & Herpin, P. (2005). Muscle fiber contractile transporter (Glut4) protein level. Biochemical Journal, 270, 397–400.
type influences the regulation of mitochondrial function. Molecular and Cellular Kim, J. M., Lee, Y. J., Choi, Y. M., Kim, B. C., Yoo, B. H., & Hong, K. C. (2008). Possible
Biochemistry, 276, 15–20. muscle fiber characteristics in the selection for improvement in porcine lean
Gueguen, N., Lefaucheur, L., Fillaut, M., Vincent, A., & Herpin, P. (2005). Control of meat production and quality. Asian-Australasian Journal of Animal Sciences, 21,
skeletal muscle mitochondria respiration by adenine nucleotides: Differential 1529–1534.
effect of ADP and ATP according to muscle contractile type in pigs. Comparative Koohmaraie, M., Kent, M. P., Shackelford, S. D., Veiseth, E., & Wheeler, T. L. (2002).
Biochemistry and Physiology Part B: Biochemistry and Molecular Biology, 140, Meat tenderness and muscle growth: Is there any relationship? Meat Science,
287–297. 62, 345–352.
Gunning, P., & Hardeman, E. (1991). Multiple mechanisms regulate muscle fiber Koohmaraie, M., Shackelford, S. D., Muggli-Cockett, N. E., & Stone, R. T. (1991). Effect
diversity. Faseb Journal, 5, 3064–3070. of the beta-adrenergic agonist L644, 969 on muscle growth, endogenous
Haddad, F., Qin, A. X., Bodell, P. W., Jiang, W., Giger, J. M., & Baldwin, K. M. (2008). proteinase activities, and postmortem proteolysis in wether lambs. Journal of
Intergenic transcription and developmental regulation of cardiac myosin heavy Animal Science, 69, 4823–4835.
chain genes. American Journal of Physiology – Heart and Circulatory Physiology, Koohmaraie, M., Shackelford, S. D., Wheeler, T. L., Lonergan, S. M., & Doumit, M. E.
294, H29–H40. (1995). A muscle hypertrophy condition in lamb (Callipyge): Characterization of
Handel, S. E., & Stickland, N. C. (1988). Catch-up growth in pigs: A relationship with effects on muscle growth and meat quality traits. Journal of Animal Science, 73,
muscle cellularity. Animal Production, 47, 291–295. 3596–3607.
Hardie, D. G., & Hawley, S. A. (2001). AMP-activated protein kinase: The energy Koulmann, N., & Bigard, A. (2006). Interaction between signalling pathways
charge hypothesis revisited. BioEssays, 23, 1112–1119. involved in skeletal muscle responses to endurance exercise. Pflugers Archiv,
Hardie, D. G., & Sakamoto, K. (2006). AMPK: A key sensor of fuel and energy status 452, 125–139.
in skeletal muscle. Physiology, 21, 48–60. Kovanen, V., Suominen, H., & Heikkinen, E. (1984). Collagen of slow twitch and fast
Harrison, A. P., Rowlerson, A. M., & Dauncey, M. J. (1996). Selective regulation of twitch muscle fibers in different types of rat skeletal muscle. European Journal of
myofibre differentiation by energy status during postnatal development. Applied Physiology, 52, 235–242.
American Journal of Physiology – Regulatory Integrative Comparative Physiology, Kurth-Kraczek, E. J., Hirshman, M. F., Goodyear, L. J., & Winder, W. W. (1999). 50
270, R667–R674. AMP-activated protein kinase activation causes GLUT4 translocation in skeletal
Hawley, S. A., Pan, D. A., Mustard, K. J., Ross, L., Bain, J., Edelman, A. M., muscle. Diabetes, 48, 1667–1671.
et al. (2005). Calmodulin-dependent protein kinase kinase-beta is an Larzul, C., Lefaucheur, L., Ecolan, P., Gogué, J., Talmant, A., Sellier, P., et al. (1997).
alternative upstream kinase for AMP-activated protein kinase. Cell Phenotypic and genetic parameters for longissimus muscle fiber characteristics
Metabolism, 2, 9–19. in relation to growth, carcass, and meat quality traits in Large White pigs.
Hawley, S. A., Selbert, M. A., Goldstein, E. G., Edelman, A. M., Carling, D., & Hardie, D. Journal of Animal Science, 75, 3126–3137.
G. (1995). 50 -AMP activates the AMP-activated protein kinase cascade, and Ca2+/ Larzul, C., Leroy, P., Gogue, J., Talmant, A., Jacquet, B., Lefaucheur, L., et al. (1999).
calmodulin activates the calmodulin-dependent protein kinase I cascade, via Selection for reduced muscle glycolytic potential in Large White pigs. II.
three independent mechanisms. Journal of Biological Chemistry, 270, Correlated responses in meat quality and muscle compositional traits. Genetics
27186–27191. Selection Evolution, 31, 61–76.
Hegarty, P. V. J., & Allen, C. E. (1978). Effects of pre-natal runting on the post-natal Lebret, B., Le Roy, P., Monin, G., Lefaucheur, L., Caritez, J. C., Talmant, A., et al. (1999).
development of skeletal muscles in swine and rats. Journal of Animal Science, 46, Influence of the three RN genotypes on chemical composition, enzyme activities
1634–1640. and myofiber characteristics of porcine skeletal muscle. Journal of Animal
Henckel, P., Oksbjerg, N., Erlandsen, E., Barton-Gade, P., & Bejerholm, C. (1997). Science, 77, 1482–1489.
Histo- and biochemical characteristics of the longissimus dorsi muscle in pigs Lefaucheur, L., Ecolan, P., Plantard, L., & Gueguen, N. (2002). New insights into
and their relationships to performance and meat quality. Meat Science, 47, muscle fiber types in the pig. Journal of Histochemistry and Cytochemistry, 50,
311–321. 719–730.
Henriksen, E. J., Bourey, R. E., Rodnick, K. J., Koranyi, L., Permutt, M. A., & Holloszy, J. Lefaucheur, L., Edom, F., Ecolan, P., & Butler-Browne, G. S. (1995). Pattern of muscle
O. (1990). Glucose transporter protein content and glucose transport capacity in fiber type formation in the pig. Developmental Dynamics, 203, 27–41.
rat skeletal muscles. American Journal of Physiology – Endocrinology and Lefaucheur, L., Hoffman, R. K., Gerrard, D. E., Okamura, C. S., Rubinstein, N., & Kelly,
Metabolism, 259, E593–E598. A. (1998). Evidence for three adult fast myosin heavy chain isoforms in type II
Herpin, P., & Lefaucheur, L. (1992). Adaptive changes in oxidative metabolism in skeletal muscle fibers in pigs. Journal of Animal Science, 76, 1584–1593.
skeletal muscle of cold-acclimated piglets. Journal of Thermal Biology, 17, Lefaucheur, L., Le Roy, P., Lebret, B., Ecolan, P., Clochefert, N., Gogué, J., et al. (2000).
277–285. Divergent selection on contractile properties of longissimus muscle fibres in the
Hochachka, P. W., & McClelland, G. B. (1997). Cellular metabolic homeostasis during pig. In Proceedings of the 46th international congress of meat science and
large-scale change in ATP turnover rates in muscles. Journal of Experimental technology, Buenos Aires (pp. 94–95). Argentina.
Biology, 100, 381–386. Lefaucheur, L., Le Dividich, J., Mourot, J., Monin, G., Ecolan, P., & Krauss, D. (1991).
Holmes, J. H. G., & Ashmore, C. R. (1972). A histochemical study of development of Influence of environmental temperature on growth, muscle and adipose tissue
muscle fiber type and size in normal and ‘‘double muscled” cattle. Growth, 36, metabolism, and meat quality in swine. Journal of Animal Science, 69,
351–372. 2844–2854.
Hurley, R. L., Anderson, K. A., Franzone, J. M., Kemp, B. E., Means, A. R., & Witters, L. Lefaucheur, L., Le Peuch, C., Barenton, B., & Vigneron, P. (1986). Characterization of
A. (2005). The Ca2+/calmodulin-dependent protein kinase kinases are AMP- insulin binding to slices of slow and fast twitch skeletal muscles in the rabbit.
activated protein kinase kinases. Journal of Biological Chemistry, 280, Hormone and Metabolic Research, 18, 725–729.
29060–29066. Lefaucheur, L., Lebret, B., Ecolan, P., Galian, M., Damon, M., Louveau, I., et al. (2008).
Jorgensen, S. B., Jensen, T. E., & Richter, E. A. (2007). Role of AMPK in skeletal muscle Sélection divergente pour la consommation alimentaire résiduelle chez le porc:
gene adaptation in relation exercise. Applied Physiology Nutrition and Effets sur les propriétés musculaires et la qualité de la viande. Journées de la
Metabolism, 32, 904–911. Recherche Porcine en France, 40, 83–84.
Jorgensen, S. B., Nielsen, J. N., Birk, J. B., Olsen, G. S., Viollet, B., Andreelli, F., et al. Lefaucheur, L., Milan, D., Ecolan, P., & Le Callennec, C. (2004). Myosin heavy chain
(2004). The alpha2-50 AMP-activated protein kinase is a site 2 glycogen synthase composition of different skeletal muscles in Large White and Meishan pigs.
kinase in skeletal muscle and is responsive to glucose loading. Diabetes, 53, Journal of Animal Science, 82, 1931–1941.
3074–3081. Lefaucheur, L., & Vigneron, P. (1986). Postnatal changes in some histochemical and
Jurie, C., Robelin, J., Picard, B., & Geay, Y. (1995). Post-natal changes in the biological enzymatic characteristics of three pig muscles. Meat Science, 16, 199–216.
characteristics of semitendinosus muscle in male Limousin cattle. Meat Science, Lengerken, G., Maak, S., Wicke, M., Fiedler, I., & Ender, K. (1994). Suitability of
41, 125–135. structural and functional traits of skeletal muscle for the genetic improvement
Kadim, I. T., Purchas, R. W., Davies, A. S., Rae, A. L., & Barton, R. A. (1993). Meat of meat quality in pigs. Archiv fur Tierzucht, 37, 133–143.
quality and muscle fibre type characteristics of southdown rams from high and Leroy, P., Elsen, J. M., Caritez, J. C., Talmant, A., Juin, H., Sellier, P., et al. (2000).
low backfat selection lines. Meat Science, 33, 97–109. Comparison between the three porcine RN genotypes for growth, carcass
Kahn, B. B., Alquier, T., Carling, D., & Hardie, D. G. (2005). AMP-activated protein composition and meat quality traits. Genetics Selection Evolution, 32, 165–186.
kinase: Ancient energy gauge provides clues to modern understanding of Leseigneur-Meynier, A., & Gandemer, G. (1991). Lipid composition of pork muscle in
metabolism. Cell Metabolism, 1, 15–25. relation to the metabolic type of the fibres. Meat Science, 29, 229–241.
Karlsson, A., Enfält, A., Essén-Gustavsson, B., Lundström, K., Rydhmer, L., & Stern, S. Lindholm, A., Persson, S., Jonsson, L., & Andrén, H. (1977). Biochemical and
(1993). Muscle histochemical and biochemical properties in relation to meat histochemical properties and fibre distribution of muscles in trained,
quality during selection for increased lean tissue growth rate in pigs. Journal of untrained and halothane sensitive pigs. Acta Agriculturae Scandinavica Section
Animal Science, 71, 930–938. A – Animal Science, 21, 357–364.
Karlsson, A. H., Klont, R. E., & Fernandez, X. (1999). Skeletal muscle fibres as factors Long, Y. C., & Zierath, J. R. (2005). Fine-tuning insulin and nitric oxide signalling by
for pork quality. Livestock Production Science, 60, 255–269. turning up the AMPs: New insights into AMP-activated protein kinase
Kauffman, R. G., Vanlaack, R. L. J. M., Russell, R. L., Pospiech, E., Cornelius, C. A., signalling. Diabetologia, 48, 2451–2453.
Suckow, C. E., et al. (1998). Can pale, soft, exudative pork be prevented by Longnus, S. L., Wambolt, R. B., Parsons, H. L., Brownsey, R. W., & Allard, M. F.
postmortem sodium bicarbonate injection? Journal of Animal Science, 76, (2003). 5-Aminoimidazole-4-carboxamide 1-beta-D-ribofuranoside (AICAR)
3010–3015. stimulates myocardial glycogenolysis by allosteric mechanisms. American
L. Lefaucheur / Meat Science 84 (2010) 257–270 269

Journal of Physiology – Regulatory Integrative and Comparative Physiology, 284, and their influences on the quantity and quality of meat from Japanese Black
R936–R944. steers. Meat Science, 54, 65–70.
Lorenzen, C. L., Koohmaraie, M., Shackelford, S. D., Jahoor, F., Freetly, H. C., Wheeler, Park, S., Scheffler, T. L., Gunawan, A. M., Shi, H., Zeng, C., Hannon, K. M., et al. (2009).
T. L., et al. (2000). Protein kinetics in callipyge lambs. Journal of Animal Science, Chronic elevated calcium blocks AMPK-induced GLUT-4 expression in skeletal
78, 78–87. muscle. American Journal of Physiology – Cell Physiology, 296, C106–C115.
Luff, A. R., & Goldspink, G. (1967). Large and small muscles. Life Sciences, 6, Pedersen, P. H., Oksbjerg, N., Karlsson, A. H., Busk, H., Bendixen, E., & Henckel, P.
1821–1826. (2001). A within litter comparison of muscle fibre characteristics and growth of
Maccatrozzo, L., Caliaro, F., Toniolo, L., Patruno, M., Reggiani, C., & Mascarello, F. halothane carrier and halothane free crossbreed pigs. Livestock Production
(2007). The sarcomeric myosin heavy chain gene family in the dog: Analysis of Science, 73, 15–24.
isoform diversity and comparison with other mammalian species. Genomics, 89, Peter, J. B., Barnard, R. J., Edgerton, V. R., Gillespie, C. A., & Stemple, K. E. (1972).
224–236. Metabolic profiles of three fiber types of skeletal muscle in guinea pigs and
Maccatrozzo, L., Patruno, M., Toniolo, L., Reggiani, C., & Mascarello, F. (2004). Myosin rabbits. Biochemistry, 11, 2627–2633.
heavy chain 2B isoform is expressed in specialized eye muscles but not in trunk Petersen, J. S., Henckel, P., Oksbjerg, N., & Sorensen, M. T. (1998). Adaptations in
and limb muscles of cattle. European Journal of Histochemistry, 48, 357–366. muscle fibre characteristics induced by physical activity in pigs. Animal Science,
Mahdavi, V., Chambers, A. P., & Nadal-Ginard, B. (1984). Cardiac alpha- and beta- 66, 733–740.
myosin heavy chain genes are organized in tandem. Proceedings of the National Pette, D., & Staron, R. S. (2000). Myosin isoforms, muscle fiber types, and transitions.
Academy of Sciences of the United States of America, 81, 2626–2630. Microscopy Research and Technique, 50, 500–509.
Mahlapuu, M., Johansson, C., Lindgren, K., Hjalm, G., Barnes, B. R., Krook, A., et al. Pette, D., & Staron, R. S. (2001). Transitions of muscle fiber phenotypic profiles.
(2004). Expression profiling of the gamma-subunit isoforms of AMP-activated Histochemistry and Cell Biology, 115, 359–372.
protein kinase suggests a major role for gamma 3 in white skeletal muscle. Picard, B., & Cassar-Malek, I. (2009). Evidence for expression of IIb myosin heavy
American Journal of Physiology – Endocrinology and Metabolism, 286, E194–E200. chain isoform in some skeletal muscles of Blonde d’Aquitaine bulls. Meat
Malenfant, P., Joanisse, D. R., Theriault, R., Goodpaster, B. H., Kelley, D. E., & Science, 82, 30–36.
Simoneau, J. A. (2001). Fat content in individual muscle fibers of lean and obese Picard, B., Lefaucheur, L., Berri, C., & Duclos, M. (2002). Muscle fibre ontogenesis in
subjects. International Journal of Obesity, 25, 1316–1321. farm animal species. Reproduction Nutrition Development, 42, 415–431.
Maltin, C. A., Sinclair, K. D., Warriss, P. D., Grant, C. M., Porter, A. D., Delday, M. I., Plaghki, L. (1985). Régénération et myogenèse du muscle strié. Journal of Physiology
et al. (1998). The effects of age at slaughter, genotype and finishing system on Paris, 80, 51–110.
the biochemical properties, muscle fibre type characteristics and eating quality Powell, S. E., & Aberle, E. D. (1981). Skeletal muscle and adipose tissue cellularity in
of bull beef from suckled calves. Animal Science, 66, 341–348. runt and normal birth weight swine. Journal of Animal Science, 52, 748–756.
Maltin, C. A., Warkup, C. C., Matthews, K. R., Grant, C. M., Porter, A. D., & Delday, M. I. Quiroz-Rothe, E., & Rivero, J. L. (2004). Coordinated expression of myosin heavy
(1997). Pig muscle fibre characteristics as a source of variation in eating quality. chains, metabolic enzymes, and morphological features of porcine skeletal
Meat Science, 47, 237–248. muscle fibre types. Microscopy Research and Technique, 65, 43–61.
Marinova, P., Lefaucheur, L., Fernandez, X., & Monin, G. (1992). Relationship Rahelic, S., & Puac, S. (1981). Fibre types in longissimus dorsi from wild and highly
between metabolism and glycogen content in skeletal muscle fibers of Large selected pig breeds. Meat Science, 5, 439–450.
White and Hampshire crossbred pigs. Journal of Muscle Foods, 3, 91–97. Rehfeldt, C., Fiedler, I., Dietl, G., & Ender, K. (2000). Myogenesis and postnatal
Marsin, A. S., Bertrand, L., Rider, M. H., Deprez, J., Beauloye, C., Vincent, M. F., et al. skeletal muscle cell growth as influenced by selection. Livestock Production
(2000). Phosphorylation and activation of heart PFK-2 by AMPK has a role in the Science, 66, 177–188.
stimulation of glycolysis during ischaemia. Current Biology, 10, 1247–1255. Rehfeldt, C., Fiedler, I., & Stickland, N. (2004). Number and size of muscle fibres in
Mascarello, F., Stecchini, M. L., Rowlerson, A., & Ballocchi, E. (1992). Tertiary relation to meat production. In M. F. W. te Pas, M. E. Everts, & H. P. Haagsman
myotubes in postnatal growing pig muscle detected by their myosin isoform (Eds.), Muscle development of livestock animals (pp. 1–38). Wallingford, UK: CAB
composition. Journal of Animal Science, 70, 1806–1813. International.
Mcallister, R. M., Reiter, B. L., Amann, J. F., & Laughlin, M. H. (1997). Skeletal muscle Rehfeldt, C., & Kuhn, G. (2006). Consequences of birth weight for postnatal growth
biochemical adaptations to exercise training in miniature swine. Journal of performance and carcass quality in pigs as related to myogenesis. Journal of
Applied Physiology, 82, 1862–1868. Animal Science, 84(Suppl. 1), E113–E123.
Meynier, A., & Gandemer, G. (1994). La flaveur des viandes cuites: Relations avec Rehfeldt, C., Tuchscherer, A., Hartung, M., & Kuhn, G. (2007). A second look at the influence
l’oxydation des phospholipides. Viandes et Produits Carnés, 15, 179–182. of birth weight on carcass and meat quality in pigs. Meat Science, 78, 170–175.
Milan, D., Jeon, J. T., Amarger, V., Robic, A., Thelander, M., Rogel-Gaillard, C., et al. Renand, G., Picard, B., Touraille, C., Berge, P., & Lepetit, J. (2001). Relationships
(2000). A mutation in PRKAG3 associated with excess glycogen content in pig between muscle characteristics and meat quality traits of young Charolais bulls.
skeletal muscle. Science, 288, 1248–1251. Meat Science, 59, 49–60.
Miller, L. R., Garwood, V. A., & Judge, M. D. (1975). Factors affecting porcine muscle Renerre, M. (1984). Variabilité entre muscles et entre animaux de la stabilité de la
fiber type, diameter and number. Journal of Animal Science, 41, 66–77. couleur des viandes bovines. Sciences des Aliments, 4, 567–584.
Miyamoto, L., Toyoda, T., Hayashi, T., Yonemitsu, S., Nakano, M., Tanaka, S., et al. Reznick, R. M., & Shulman, G. I. (2006). The role of AMP-activated protein kinase in
(2007). Effect of acute activation of 50 -AMP-activated protein kinase on mitochondrial biogenesis. Journal of Physiology, 574, 33–39.
glycogen regulation in isolated rat skeletal muscle. Journal of Applied Rinaldi, C., Haddad, F., Bodell, P. W., Qin, A. X., Jiang, W. H., & Baldwin, K. M. (2008).
Physiology, 102, 1007–1013. Intergenic bidirectional promoter and cooperative regulation of the IIx and IIb
Monin, G., Mejenes-Quijano, A., Talmant, A., & Sellier, P. (1987). Influence of breed MHC genes in fast skeletal muscle. American Journal of Physiology – Regulatory
and muscle metabolic type on muscle glycolytic potential and meat pH in pigs. Integrative and Comparative Physiology, 295, R208–R218.
Meat Science, 20, 149–158. Rinaldo, D., & Le Dividich, J. (1991). Effects of warm exposure on adipose tissue and
Naya, F. J., Mercer, B., Shelton, J., Richardson, J. A., Williams, R. S., & Olson, E. N. muscle metabolism in growing pigs. Comparative Biochemistry and Physiology
(2000). Stimulation of slow skeletal muscle fiber gene expression by calcineurin Part A, 100A, 995–1002.
in vivo. Journal of Biological Chemistry, 275, 4545–4548. Rincker, P. J., Killefer, J., Ellis, M., Brewer, M. S., & Mckeith, F. K. (2008).
Ngapo, T., Berge, P., Culioli, J., Dransfield, E., de Smet, S., & Claeys, E. (2002). Intramuscular fat content has little influence on the eating quality of fresh
Perimysial collagen crosslinking and meat tenderness in Belgian Blue double- pork loin chops. Journal of Animal Science, 86, 730–737.
muscled cattle. Meat Science, 61, 91–102. Rincker, P. J., Killefer, J., Matzat, P. D., Carr, S. N., & Mckeith, F. K. (2009). The effect of
Ngapo, T. M., & Gariepy, C. (2008). Factors affecting the eating quality of pork. ractopamine and intramuscular fat content on sensory attributes of pork from
Critical Reviews in Food Science and Nutrition, 48, 599–633. pigs of similar genetics. Journal of Muscle Foods, 20, 79–88.
Nii, M., Hayashi, T., Mikawa, S., Tani, F., Niki, A., Mori, N., et al. (2005). Quantitative Rivero, J. L. L., Serrano, A. L., Barrey, E., Valette, J. P., & Jouglin, M. (1999). Analysis of
trait loci mapping for meat quality and muscle fiber traits in a Japanese wild myosin heavy chains at the protein level in horse skeletal muscle. Journal of
boar  Large White intercross. Journal of Animal Science, 83, 308–315. Muscle Research and Cell Motility, 20, 211–221.
Oksbjerg, N., Henckel, P., Andersen, S., Pedersen, B., & Nielsen, B. (2004). Genetic Rosenvold, K., & Andersen, H. J. (2003). Factors of significance, for pork quality – A
variation of in vivo muscle glycerol, glycogen, and pigment in Danish purebred review. Meat Science, 64, 219–237.
pigs. Acta Agriculturae Scandinavica Section A: Animal Science, 54, 187–192. Rosenvold, K., Petersen, J. S., Laerke, H. S., Jensen, S. K., Therkildsen, M., Karlsson, A.
Oksbjerg, N., Petersen, J. S., Sorensen, I. L., Henckel, P., Vestergaard, M., Ertbjerg, P., H., et al. (2001). Muscle glycogen stores and meat quality as affected by
et al. (2000). Long-term changes in performance and meat quality of Danish strategic finishing feeding of slaughter pigs. Journal of Animal Science, 79,
Landrace pigs: A study on a current compared with an unimproved genotype. 382–391.
Animal Science, 71, 81–92. Ross, J. J., Duxson, M. J., & Harris, A. J. (1987). Formation of primary and secondary
Ontell, M., Bourke, D., & Hughes, D. (1988). Cytoarchitecture of the fetal murine myotubes in rat lumbrical muscles. Development, 100, 383–394.
soleus muscle. American Journal of Anatomy, 181, 267–278. Ruusunen, M., & Puolanne, E. (2004). Histochemical properties of fibre types in
Ouali, A. (1990). Meat tenderization: Possible causes and mechanisms. A review. muscles of wild and domestic pigs and the effect of growth rate on muscle fibre
Journal of Muscle Foods, 1, 129–165. properties. Meat Science, 67, 533–539.
Ouali, A., & Talmant, A. (1990). Calpains and calpastatin distribution in bovine, Ruusunen, M., Sevonaimonen, M. L., & Puolanne, E. (1996). Composition and cross-
porcine and ovine skeletal muscles. Meat Science, 28, 331–348. sectional area of muscle fibre types in relation to daily gain and lean and fat
Ouhayoun, J., & Beaumont, A. (1968). Etude du caractère culard: Anatomie content of carcass in Landrace and Yorkshire pigs. Agricultural and Food Science
microscopique comparée du tissu musculaire de mâle Charolais normaux et in Finland, 5, 593–600.
culards. Annales de Zootechnie, 17, 213–223. Ryu, Y. C., Choi, Y. M., Ko, Y., & Kim, B. C. (2007). Relationship between serum
Ozawa, S., Mitsuhashi, T., Mitsumoto, M., Matsumoto, S., Itoh, N., Itagaki, K., et al. endocrine factors, histochemical characteristics of Longissimus dorsi muscle
(1999). The characteristics of muscle fiber types of longissimus thoracis muscle and meat quality in pigs. Journal of Muscle Foods, 18, 95–108.
270 L. Lefaucheur / Meat Science 84 (2010) 257–270

Ryu, Y. C., Choi, Y. M., Lee, S. H., Shin, H. G., Choe, J. H., Kim, J. M., et al. (2008). Toniolo, L., Maccatrozzo, L., Patruno, M., Pavan, E., Caliaro, F., Rossi, R., et al. (2007).
Comparing the histochemical characteristics and meat quality traits of different Fiber types in canine muscles: Myosin isoform expression and functional
pig breeds. Meat Science, 80, 363–369. characterization. American Journal of Physiology – Cell Physiology, 292,
Ryu, Y. C., & Kim, B. C. (2005). The relationship between muscle fiber characteristics, C1915–C1926.
postmortem metabolic rate, and meat quality of pig longissimus dorsi muscle. Totland, G. K., Kryvi, H., & Slinde, E. (1988). Composition of muscle fibre types and
Meat Science, 71, 351–357. connective tissue in bovine M. Semitendinosus and its relation to tenderness.
Ryu, Y. C., & Kim, B. C. (2006). Comparison of histochemical characteristics in Meat Science, 23, 303–315.
various pork groups categorized by postmortem metabolic rate and pork Tribout, T., Caritez, J. C., Cogué, J., Gruand, J., Bouffaud, M., Billon, Y., et al. (2004).
quality. Journal of Animal Science, 84, 894–901. Estimation, par utilisation de semence congelée, du progrès génétique réalisé en
Ryu, Y. C., Lee, M. H., Lee, S. K., & Kim, B. C. (2006). Effects of muscle mass and fiber France entre 1977 et 1998 dans la race porcine Large White: Résultats pour
type composition of Longissimus dorsi muscle on postmortem metabolic rate quelques caractères de production et de qualité des tissus gras et maigres.
and meat quality in pigs. Journal of Muscle Foods, 17, 343–353. Journées de la Recherche Porcine en France, 36, 275–282.
Ryu, Y. C., Rhee, M. S., & Kim, B. C. (2004). Estimation of correlation coefficients Turinsky, J., O’Sullivan, D. M., & Bayly, B. P. (1990). 1,2-Diacylglycerol and ceramide
between histological parameters and carcass traits of pig Longissimus dorsi levels in insulin-resistant tissues of the rat in vivo. Journal of Biological
muscle. Asian-Australasian Journal of Animal Sciences, 17, 428–433. Chemistry, 265, 16880–16885.
Scheffler, T. L., & Gerrard, D. E. (2007). Mechanisms controlling pork quality Underwood, K. R., Means, W. J., Zhu, M. J., Ford, S. P., Hess, B. W., & Du, M. (2008).
development: The biochemistry controlling postmortem energy metabolism. AMP-activated protein kinase is negatively associated with intramuscular fat
Meat Science, 77, 7–16. content in longissimus dorsi muscle of beef cattle. Meat Science, 79, 394–402.
Schiaffino, S., Gorza, L., Sartore, S., Saggin, L., Ausoni, S., Vianello, M., et al. (1989). Underwood, K. R., Tong, J., Zhu, M. J., Shen, Q. W., Means, W. J., Ford, S. P., et al.
Three myosin heavy chain isoforms in type 2 skeletal muscle fibres. Journal of (2007). Relationship between kinase phosphorylation, muscle fiber typing, and
Muscle Research and Cell Motility, 10, 197–205. glycogen accumulation in Longissimus muscle of beef cattle with high and low
Schiaffino, S., & Reggiani, C. (1994). Myosin isoforms in mammalian skeletal muscle. intramuscular fat. Journal of Agricultural and Food Chemistry, 55, 9698–9703.
Journal of Applied Physiology, 77, 493–501. Valin, C., Touraille, C., Vigneron, P., & Ashmore, C. R. (1982). Prediction of lamb meat
Schiaffino, S., & Reggiani, C. (1996). Molecular diversity of myofibrillar proteins: quality traits based on muscle biopsy fibre typing. Meat Science, 6, 257–263.
Gene regulation and functional significance. Physiological Reviews, 76, 371–423. Van den Maagdenberg, K., Stinckens, A., Lefaucheur, L., Buys, N., & de Smet, S.
Seideman, S. C., & Crouse, J. D. (1986). The effects of sex condition, genotype and (2008). The effect of mutations in the insulin-like growth factor-II and
diet on bovine muscle fiber characteristics. Meat Science, 17, 55–72. ryanodine receptor-1 genes on biochemical and histochemical muscle fibre
Seideman, S. C., Crouse, J. D., & Cross, H. R. (1986). The effect of sex condition and characteristics in pigs. Meat Science, 79, 757–766.
growth implants on bovine muscle fiber characteristics. Meat Science, 17, 79–95. van Loon, L. J. C., & Goodpaster, B. H. (2006). Increased intramuscular lipid storage in
Seideman, S. C., Crouse, J. D., & Mersmann, H. J. (1989). Carcass, muscle and meat the insulin-resistant and endurance-trained state. Pflugers Archiv, 451, 606–616.
characteristics of lean and obese pigs. Journal of Animal Science, 67, 2950– Villa-Moruzzi, E., Locci-Cubeddu, T., & Bergamini, E. (1979). Developmental changes
2955. of glycogen enzymes in fast and slow muscles of the rat. Growth, 43, 73–79.
Sellier, P. (1998). Genetics of meat and carcass traits. In M. F. Rothschild & A. Vuocolo, T., Byrne, K., White, J., McWilliam, S., Reverter, A., Cockett, N. E., et al.
Ruvinsky (Eds.), Genetics of the pig (pp. 463–510). Wallingford, UK: CAB (2007). Identification of a gene network contributing to hypertrophy in
International. callipyge skeletal muscle. Physiological Genomics, 28, 253–272.
Serrano, A. L., Perez, M., Lucia, A., Chicharro, J. L., QuirozRothe, E., & Rivero, J. L. L. Wegner, J., Albrecht, E., Fiedler, I., Teuscher, F., Papstein, H. J., & Ender, K. (2000).
(2001). Immunolabelling, histochemistry and in situ hybridisation in human Growth-and breed-related changes of muscle fiber characteristics in cattle.
skeletal muscle fibres to detect myosin heavy chain expression at the protein Journal of Animal Science, 78, 1485–1496.
and mRNA level. Journal of Anatomy, 199, 329–337. Weiler, U., Appell, H. J., Kremser, M., Hofäcker, S., & Claus, R. (1995). Consequences
Shen, Q. W., Gerrard, D. E., & Du, M. (2008). Compound C, an inhibitor of AMP- of selection on muscle composition. A comparative study on gracilis muscle in
activated protein kinase, inhibits glycolysis in mouse longissimus dorsi wild and domestic pigs. Anatomia Histologia Embryologia, 24, 77–80.
postmortem. Meat Science, 78, 323–330. Weiss, A., Mcdonough, D., Wertman, B., Acakposatchivi, L., Montgomery, K.,
Shrager, J. B., Desjardins, P. R., Burkman, J. M., Konig, S. K., Stewart, D. R., Su, L., et al. Kucherlapati, T., et al. (1999). Organization of human and mouse skeletal
(2000). Human skeletal myosin heavy chain genes are tightly linked in the order myosin heavy chain gene clusters is highly conserved. Proceedings of the
embryonic-IIa–IId/x-IIb-perinatal-extraocular. Journal of Muscle Research and National Academy of Sciences of the United States of America, 96, 2958–2963.
Cell Motility, 21, 345–355. Weiss, A., Schiaffino, S., & Leinwand, L. A. (1999). Comparative sequence analysis of
Smerdu, V., Karsch-Mizrachi, I., Campione, M., Leinwand, L., & Schiaffino, S. (1994). the complete human sarcomeric myosin heavy chain family: Implications for
Type IIx myosin heavy chain transcripts are expressed in type IIb fibers of functional diversity. Journal of Molecular Biology, 290, 61–75.
human skeletal muscle. American Journal of Physiology – Cell Physiology, 267, West, R. L. (1974). Red to white fibre ratios as an index of double muscling in beef
C1723–C1728. cattle. Journal of Animal Science, 38, 1165–1175.
Smerdu, V., Strbenc, M., Meznaric-Petrusa, M., & Fazarinc, G. (2005). Identification Wicke, M., Von Lengerken, G., Fiedler, I., Altmann, M., & Ender, K. (1991). Influence
of myosin heavy chain I, IIa and IIx in canine skeletal muscles by an of selection based on muscle structure characteristics of the M. Longissimus on
electrophoretic and immunoblotting study. Cells Tissues Organs, 180, 106–116. stress sensitivity and meat quality in the pig. Fleischwirtschaft, 71, 437–442.
Solomon, M. B., & Lynch, G. P. (1988). Biochemical, histochemical and palatability Wigmore, P. M., & Stickland, N. C. (1983). Muscle development in large and small
characteristics of young ram lambs as affected by diet and electrical pig fetuses. Journal of Anatomy, 137, 235–245.
stimulation. Journal of Animal Science, 66, 1955–1962. Wimmers, K., Murani, E., Ngu, N. T., Schellander, K., & Ponsuksili, S. (2007).
Solomon, M. B., & West, R. L. (1985). Profile of fiber types in muscles from wild pigs Structural and functional genomics to elucidate the genetic background of
native to the United States. Meat Science, 13, 247–254. microstructural and biophysical muscle properties in the pig. Journal of Animal
Stickland, N. C., & Handel, S. E. (1986). The numbers and types of muscle fibers in Breeding and Genetics, 124, 27–34.
large and small breeds of pigs. Journal of Anatomy, 147, 181–189. Wimmers, K., Ngu, N. T., Jennen, D. G. J., Tesfaye, D., Murani, E., Schellander, K., et al.
Suzuki, K., Irie, M., Kadowaki, H., Shibata, T., Kumagai, M., & Nishida, A. (2005). (2008). Relationship between myosin heavy chain isoform expression and
Genetic parameter estimates of meat quality traits in Duroc pigs selected for muscling in several diverse pig breeds. Journal of Animal Science, 86, 795–803.
average daily gain, longissimus muscle area, backfat thickness, and Wojtaszewski, J. F. P., Jorgensen, S. B., Hellsten, Y., Hardie, D. G., & Richter, E. A.
intramuscular fat content. Journal of Animal Science, 83, 2058–2065. (2002). Glycogen-dependent effects of 5-aminoimidazole-4-carboxamide
Swatland, H. J. (1984). The radial distribution of succinate dehydrogenase activity in (AICA)-riboside on AMP-activated protein kinase and glycogen synthase
porcine muscle fibres. Histochemical Journal, 16, 329–381. activities in rat skeletal muscle. Diabetes, 51, 284–292.
Swatland, H. J., & Kieffer, N. M. (1974). Fetal development of the double muscled Woods, A., Dickerson, K., Heath, R., Hong, S. P., Momcilovic, M., Johnstone, S. R., et al.
condition in cattle. Journal of Animal Science, 38, 752–757. (2005). Ca2+/calmodulin-dependent protein kinase kinase-beta acts upstream of
Swynghedauw, B. (1986). Developmental and functional adaptation of contractile AMP-activated protein kinase in mammalian cells. Cell Metabolism, 2, 21–33.
proteins in cardiac and skeletal muscles. Physiological Reviews, 66, 710–771. Wright, D. C., Geiger, P. C., Holloszy, J. O., & Ho-Han, D. (2005). Contraction and
Tanabe, R., Muroya, S., & Chikuni, K. (1998). Sequencing of the 2a, 2x, and slow hypoxia stimulated glucose transport is mediated by a Ca2+ dependent
isoforms of the bovine myosin heavy chain and the different expression among mechanism in slow twitch rat soleus muscle. American Journal of Physiology –
muscles. Mammalian Genome, 9, 1056–1058. Endocrinology and Metabolism, 288, E1062–E1066.
Tate, C. A., Hyek, M. F., & Taffet, G. E. (1991). The role of calcium in the energetics of Yu, H. Y., Fujii, N., Hirshman, M. F., Pomerleau, J. M., & Goodyear, L. J. (2004). Cloning
contracting skeletal muscle. Sports Medicine, 12, 208–217. and characterization of mouse 50 -AMP-activated protein kinase gamma 3
Toniolo, L., Maccatrozzo, L., Patruno, M., Caliaro, F., Mascarello, F., & Reggiani, C. subunit. American Journal of Physiology – Cell Physiology, 286, C283–C292.
(2005). Expression of eight distinct MHC isoforms in bovine striated muscles: Zamora, F., Debiton, E., Lepetit, J., Lebert, A., Dransfield, E., & Ouali, A. (1996).
Evidence for MHC-2B presence only in extraocular muscles. Journal of Predicting variability of ageing and toughness in beef M. Longissimus
Experimental Biology, 208, 4243–4253. lumborum et thoracis. Meat Science, 43, 321–333.

You might also like