Molecular and biochemical regulation of skeletal muscle metabolism
Copyright © The Author(s) 2022. Published by Oxford University Press on behalf of the American Society of Animal Science. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com.
This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model)
Abstract
Skeletal muscle hypertrophy is a culmination of catabolic and anabolic processes that are interwoven into major metabolic pathways, and as such modulation of skeletal muscle metabolism may have implications on animal growth efficiency. Muscle is composed of a heterogeneous population of muscle fibers that can be classified by metabolism (oxidative or glycolytic) and contractile speed (slow or fast). Although slow fibers (type I) rely heavily on oxidative metabolism, presumably to fuel long or continuous bouts of work, fast fibers (type IIa, IIx, and IIb) vary in their metabolic capability and can range from having a high oxidative capacity to a high glycolytic capacity. The plasticity of muscle permits continuous adaptations to changing intrinsic and extrinsic stimuli that can shift the classification of muscle fibers, which has implications on fiber size, nutrient utilization, and protein turnover rate. The purpose of this paper is to summarize the major metabolic pathways in skeletal muscle and the associated regulatory pathways.
Keywords: muscle, metabolism, nutrients, satellite cells
This review summarizes the major metabolic pathways in skeletal muscle and the associated regulatory pathways.
Introduction
Beyond its fundamental function of locomotion and thermogenesis, skeletal muscle plays an essential role in whole body metabolic equilibrium. In healthy individuals, skeletal muscle represents about 40% of body mass, is a significant contributor to basal energy expenditure (Zurlo et al., 1990), and accounts for 70%–90% of insulin-mediated glucose disposal (DeFronzo et al., 1981; Shulman et al., 1990). As a result, modulation of skeletal muscle metabolism and insulin sensitivity has proven to be an effective means of alleviating metabolic disorders in humans (Izumiya et al., 2008; Meng et al., 2013; Song et al., 2013). Skeletal muscles are composed of a heterogeneous population of muscle cells with distinct metabolic characteristics (Holloszy, 1967) and demonstrate enormous plasticity to adapt to changing external stimuli including activity, load, nutritional factors, and external conditions. For example, endurance exercise or resistance training increases metabolic demand and subsequently increases skeletal muscle mitochondrial content and oxidative capacity (Donges et al., 2012; Di Donato et al., 2014). Contrarily, conditions associated with the onset of obesity such as excessive calorie intake, in combination with a reduction in workload, decrease oxidative capacity (Mootha et al., 2003), which is associated with metabolic syndrome and poor insulin sensitivity (Simoneau et al., 1995). However, diabetic and obese patients subjected to exercise training regimes have improved insulin sensitivity, albeit significantly blunted compared to healthy individuals (Rosholt et al., 1994; Han et al., 1997; Fueger et al., 2004). Overall, greater muscle oxidative capacity can provide protection against metabolic disorders, attenuate age-related muscle atrophy, and alleviate metabolic defects associated with some myopathies (Conley et al., 2000; Wang et al., 2004; Wenz et al., 2008; Wenz et al., 2009; Ljubicic et al., 2011). Although skeletal muscle metabolism has become the center of attention in the fight against metabolic disorders in humans, modulation of skeletal muscle metabolism in meat-producing animals as a means to improve production efficiency remains largely unexplored.
Although it is apparent that glycolytic metabolism in skeletal muscle is associated with more efficient lean growth, the regulatory mechanisms that drive metabolic changes are undefined. For example, a greater propensity for glycolytic metabolism is associated with high feed efficiency (FE) in pigs, whereas low FE pigs exhibit an increase in oxidative enzyme expression (Le Naou et al., 2012; Fu et al., 2017). Furthermore, ractopamine, a pharmacological agent that promotes lean accretion and improves FE (Williams et al., 1994; Main et al., 2009; Hinson et al., 2011), induces a shift in muscle fiber composition to increase glycolytic capacity (Depreux et al., 2002; Gunawan et al., 2007; Gunawan et al., 2020). However, the adaptive mechanism that facilitates these changes remains elusive. Herein, we review the basics of skeletal muscle metabolism and then delve into our current understanding of skeletal muscle metabolic regulation.
The biochemistry of skeletal muscle metabolism
Skeletal muscle has the remarkable ability to adjust its energy expenditure according to energy demands. Within milliseconds, muscle can increase its energy expenditure 100-fold to transition between resting and active states (Sahlin et al., 1998). Despite these large fluctuations in energy demands, adenosine triphosphate (ATP) concentrations remain relatively constant (Arnold et al., 1984; Heineman and Balaban, 1990), which depicts the striking precision of cellular energy production systems to equal demand. This metabolic control is governed by intricate signaling and feedback loops that relay cellular energy status to alter metabolic pathways within the cell.
ATP homeostasis
Skeletal muscle metabolism is a series of complex biochemical processes that convert nutrients into chemical energy in the form of adenosine triphosphate (ATP; Figure 1 ). There are several metabolic pathways found in muscle cells that have different subcellular locations and substrate utilization with varying longevity. For brief, rapid mobilization of ATP, the phosphagen system is a reversible reaction that catalyzes the degradation of phosphocreatine (PCr) by creatine kinase to produce creatine (Cr) and brief supplies of ATP from adenosine diphosphate (ADP) (PCr + ADP ↔ Cr + ATP). Additionally, the enzyme adenylate kinase collaborates with the phosphagen system to contribute to short-term energy buffering to produce 1 ATP and 1 adenosine monophosphate (AMP) from 2 ADP (2 ADP ↔ 1 ATP + 1 AMP). Adenylate kinase is also an important contributor to reporting cellular energy status by translating relatively small changes in the ATP:ADP ratio into large changes in AMP concentrations. The concentration of cellular AMP is a sensitive indicator of cellular energy status, and small increases are potent allosteric activators of several metabolic enzymes to increase ATP production. For example, 5ʹ-adenosine monophosphate-activated protein kinase (AMPK) increases glycogenolysis by activating glycogen phosphorylase and increases glucose uptake through greater glucose transporter type 4 (GLUT4) expression (Zheng et al., 2001) and translocation (Kurth-Kraczek et al., 1999). In fact, chronic activation of AMPK alters skeletal muscle metabolic characteristics. Furthermore, AMPK activation increases glucose conversation to glucose-6-phosphate through hexokinase activation (Stoppani et al., 2002), the first step of glycolysis. These temporary energy buffering systems maintain ATP levels until other energy production systems are fully active.
Major metabolic pathways in skeletal muscle. Glucose or glycogen can be metabolized through glycolysis to yield pyruvate, which can be converted to acetyl-CoA by PDH or converted to lactate by LDH. Alternatively, the glycolytic intermediates G6P or F6P can be redirected to the PPP, for biosynthetic processes, or to the HBP, for metabolic signaling. If pyruvate is converted to acetyl-CoA, it can feed into the TCA cycle to be fully oxidized, to produce the reducing equivalents NADH and FADH2, or exit the cycle to contribute to biosynthetic processes, such as fatty acid or amino acid synthesis. Created with BioRender.com.
Glycolysis
Glycolysis converts glucose into pyruvate through a series of enzymatic reactions and can produce ATP 100 times faster than oxidative phosphorylation (OxPhos) (Pfeiffer et al., 2001). Large glycogen stores in muscle allow for rapid mobilization of glucose-1-phosphate by glycogen phosphorylase to fuel glycolysis, which is the predominate energy source during high-intensity exercise (Katz et al., 1986). Alternatively, GLUT4 permits plasma glucose uptake, although rate of uptake is limited by the transport of GLUT4 to the cell surface (Zisman et al., 2000). Translocation of GLUT4 vesicles can be stimulated by insulin in response to high blood glucose levels after feeding; however, its translocation is impaired with insulin resistance and/or diabetes (Zorzano et al., 1996; Zierath et al., 1998; Tremblay et al., 2001). Alternatively, contractile activity from acute exercise can stimulate GLUT4 translocation in an insulin-independent manner (Hayashi et al., 1997; Kennedy et al., 1999; Jessen and Goodyear, 2005). Once within the cells, glucose is rapidly phosphorylated by hexokinase to produce glucose-6-phosphate (G6P) to prevent its outward diffusion and maintain the concentration gradient for continued glucose influx. In either case of glycogen breakdown or glucose uptake, G6P can then branch into anabolic (pentose phosphate pathway), metabolic signaling (hexosamine biosynthetic pathway), or catabolic (glycolysis) pathways depending on cellular demands.
Pentose phosphate pathway
The pentose phosphate pathway (PPP) emanates from the glycolytic intermediates G6P and fructose-6-phosphate (F6P) for synthesis of metabolites that are necessary for skeletal muscle anabolism including ribulose-5-phosphate for nucleic acid synthesis, erythrose-4-phosphate for aromatic amino acid synthesis, and nicotinamide adenine dinucleotide phosphate (NADPH) for reductive biosynthesis reactions such as lipogenesis. The two enzymes glucose-6-phosphate dehydrogenase (G6PDH) and 6-phosphogluconate dehydrogenase (6PGDH) are often used to assess PPP activity because they catalyze the first committed step in the PPP and the final step that produces ribulose 5-phosphate, respectively. Activity of the PPP is quite low in skeletal muscle compared to other tissue types (Battistuzzi et al., 1985), which is not surprising considering skeletal muscle is a terminally differentiated tissue. However, in regenerating muscle, 24 h after damage, G6PDH and 6PGDH activities increase 9-fold compared to the undamaged controls (Wagner et al., 1978). Furthermore, skeletal muscle-specific overexpression of AKT serine/threonine kinase 1 (AKT1) induces muscle hypertrophy (Izumiya et al., 2008), accompanied by a 3.5- and 2.3-fold increase in G6PDH and 6PGDH transcripts, respectively (Wu et al., 2017). These results collectively suggest greater PPP flux augments skeletal muscle growth.
Hexosamine biosynthetic pathway
The hexosamine biosynthetic pathway (HBP) is a metabolic signaling pathway that diverts about 2%–3% of cellular glucose from glycolysis (Marshall et al., 1991). It originates from the glycolytic intermediate F6P and is catalyzed by the rate-limiting enzyme glutamine: fructose 6-phosphate amidotransferase (GFAT), which converts F6P and glutamine to glucosamine-6-phosphate and glutamate. The subsequent steps of the HBP generate uridine diphosphate N-acetylglucosamine (UDP-GlcNAc). Collectively, UDP-GlcNAc is indicative of the cellular nutrient flux by integrating carbohydrate, lipid, nucleotide, and protein metabolism (Slawson et al., 2010; Hanover et al., 2012; Ruan et al., 2013). As such, the level of UDP-GlcNAc is an indicator of cellular nutrient status, as it reflects the conditions of nutrient excess or scarcity. Ultimately, UDP-GlcNAc is a donor substrate for the post-translational modification, O-linked-N-acetylglucosaminylation (O-GlcNAcylation), which modifies cell behavior according to fluctuating nutrient conditions (Harwood and Hanover, 2014).
The TCA cycle and oxidative phosphorylation
On the other hand, once F6P is converted to fructose-1,6-bisphosphate by phosphofructokinase (PFK), it is committed to glycolysis and then undergoes a sequential, multi-step conversion to pyruvate. Pyruvate can then be converted to either lactate by lactate dehydrogenase (LDH) or acetyl CoA by pyruvate dehydrogenase (PDH). During periods of rapid glycolytic flux, or in anaerobic conditions, pyruvate is converted to lactate which can then be shuttled out of the cell through monocarboxylase transporters (MCT) to the liver for hepatic gluconeogenesis to replenish glycogen stores (Brooks, 1986). Alternatively, pyruvate can be transported into the mitochondrial matrix for oxidation but must first traverse the outer mitochondrial membrane, presumably through the porin channel (Huizing et al., 1996), and the inner mitochondrial membrane, through the mitochondrial pyruvate carrier (Bricker et al., 2012; Herzig et al., 2012). Once inside the matrix, PDH converts pyruvate into acetyl-CoA to enter into the tricarboxylic acid (TCA) cycle, or less frequently, pyruvate carboxylase converts pyruvate into oxaloacetate to support anaplerosis. As PDH is the major link between glycolysis and the TCA cycle, it is an important molecular switch that controls the transition between the two metabolic pathways and is tightly regulated via phosphorylation. A balance between pyruvate dehydrogenase kinases (PDK) and pyruvate dehydrogenase phosphatases (PDP) regulates the pyruvate dehydrogenase complex (PDC) in response to nutritional status. Pyruvate dehydrogenase kinase 4 (PDK4) is largely found in skeletal muscle (Bowker-Kinley et al., 1998) and has an increased activity during periods of starvation or high fat feeding (Wu et al., 1998; Holness et al., 2000; Jeoung et al., 2006; Jeoung and Harris, 2008), which suggests that PDK4 plays an important role in dietary regulation of carbohydrate oxidation.
Fatty acids also fuel the TCA cycle through β-oxidation, which breaks down fatty acid chains by cleaving two carbons per cycle to produce acetyl-CoA. The four main enzymes involved in the β-oxidation cycle are acyl-CoA dehydrogenase, enoyl-CoA hydratase, hydroxyacyl-CoA dehydrogenase, and ketoacyl-CoA thiolase. One cycle produces acetyl-CoA, a fatty acid chain that is two carbons shorter, nicotinamide adenine dinucleotide (NADH), and flavin adenine dinucleotide (FADH2). Acetyl-CoA then proceeds into the TCA cycle and the reducing equivalents are oxidized in the electron transport chain (ETC) to produce ATP. Products of these reactions are allosteric regulators of β-oxidation enzymes including their intermediate product, the ratio of NADH/NAD+, and acetyl-CoA levels (Eaton, 2002). Peroxisome proliferator-activated receptors (PPARs) and coactivator peroxisome proliferator-activated receptor-gamma coactivator 1α (PGC1α) are the major transcriptional regulators of β-oxidation. In fact, deletion of PPARβ in skeletal muscle results in a metabolic shift that decreases muscle oxidative capacity (Schuler et al., 2006) and ablation of skeletal muscle PGC1α increases the number of glycolytic fibers (Handschin et al., 2007), which may have implications on feed efficiency. Indeed, pigs with improved measurements of feed efficiency downregulate mitochondrial energy metabolic enzymes (Jing et al., 2015; Fu et al., 2017).
Acetyl-CoA can be fully oxidized in the TCA cycle to produce the reducing equivalents NADH and FADH2 for ATP synthesis, or its intermediates can exit the cycle through biosynthetic processes. Although the TCA cycle is often referred to as a “cycle,” implying a unidirectional and defined series of reactions, there are many entry and exit points that allow for both catabolic and anabolic processes in response to different cellular demands. During anabolic conditions, TCA intermediates can be shuttled from the mitochondrial matrix into the cytosol for fatty acid, amino acid, and nucleotide synthesis (Ahn and Metallo, 2015). On the other hand, high energy demand elicits catabolic processes, and intermediates undergo multiple oxidation steps to supply the electron transport chain (ETC) with reducing equivalents (NADH and FADH2) to replenish ATP stores through OxPhos.
Not surprisingly, the TCA cycle is tightly coupled to energy demand through complex feedback loops. For example, accumulation of NADH is a potent inhibitor of all TCA cycle regulatory enzymes, and as such, malfunctions in the ETC have implications on the TCA cycle (Martinez-Reyes and Chandel, 2020). Similarly, high ATP concentrations or accumulation of TCA cycle intermediates inhibit several TCA enzymes. On the other hand, high energy demands increase AMP levels, which is an allosteric activator of the regulatory TCA cycle enzymes. Although allosteric regulation provides dynamic feedback to keep the TCA cycle in-check, it is also governed by other signaling pathways that provide another layer of regulation. In skeletal muscle, mitochondria are largely found anchored to the sarcoplasmic reticulum (SR) (Boncompagni et al., 2009), which allows for bidirectional signaling between the two organellar systems. More specifically, the calcium released from the SR during muscle contraction can be sequestered by mitochondria, which subsequently stimulates the TCA cycle and mitochondrial ATP production by activating pyruvate dehydrogenase, α-ketoglutarate dehydrogenase, isocitrate dehydrogenase, and the F1F0-ATPase (McCormack and Denton, 1979; McCormack et al., 1990; McCormack and Denton, 1993; Hansford, 1994). The parallel activation of muscle contraction and mitochondrial metabolism simultaneously initiates muscle contraction and ATP production to match energy demand with energy supply effectively (Korzeniewski, 2007). In summary, skeletal muscle metabolism is tightly regulated through a series of complex biochemical processes orchestrated by a hierarchy of catabolic and anabolic events.
Mitochondria bioenergetics and dynamics
Multifaceted functions of mitochondria
Although mitochondria are well-known for ATP production, they also play a central role in metabolic signaling (Hansford and Zorov, 1998; Bohovych and Khalimonchuk, 2016; Martinez-Reyes and Chandel, 2020), reactive oxygen species (ROS) production/signaling and scavenging (Skulachev, 1996; Aon et al., 2010; Brand, 2010), cell death (Goldstein et al., 2000), ion and metabolite transport (Palmieri and Pierri, 2010), iron-sulfur protein biogenesis (Nuth et al., 2002; Bulteau et al., 2004; Yoon and Cowan, 2004), and calcium homeostasis (Deluca and Engstrom, 1961; Vasington and Murphy, 1962; Rottenberg and Scarpa, 1974). Mitochondria have two membranes, the porous outer mitochondrial membrane (OMM) and selective inner mitochondrial membrane (IMM), which give rise to the inner membrane space (IMS) and mitochondrial matrix. Residing in the IMM, electron transport chain complexes transfer electrons from reducing equivalents to electron acceptors through several redox reactions that are coupled with the transfer of protons from the matrix to the IMS, generating an electrochemical gradient ( Figures 1 and and2). 2 ). The mitochondrial membrane potential, or electrochemical gradient, is harnessed for ATP production but is also the driving force for its other central metabolic roles.
The electron transport chain and generation of the electrochemical gradient. The ETC consists of four protein complexes (CI-CIV) and two electron transfer carriers, ubiquinone and cytochrome c. The electron donors, NADH and FADH2, are oxidized by CI or CII, respectively, to reduce the first electron carrier ubiquinone to ubiquinol. Although both CI and CII transport electrons, CI also pumps four protons into the IMS, whereas CII does not pump protons. CIII catalyzes the transfer of electrons from the first electron carrier to the second (ubiquinol to cytochrome c) and pumps four protons into the IMS. Complex IV catalyzes the transfer of electrons from cytochrome c to the terminal electron carrier oxygen to produce water and pumps protons. Complexes are arranged into super complexes in the cristae, often into arrangements of I + III2 + IV, III2 + IV, and dimers of ATP synthase (CV). In normal physiological conditions, proton re-entry into the matrix largely occurs through the ATP synthase (CV), coupling the electrochemical gradient to ATP production. However, there are alternative re-entry points that are uncoupled from ATP production, such as uncoupling proteins (UCPs) to produce heat and the nicotinamide nucleotide transhydrogenase (NNT) to replenish NAPDH pools for biosynthetic processes and ROS scavenging. Created with BioRender.com.
Mitochondria in different subcellular locations also have variable roles in skeletal muscle physiology. Intrafibrillar mitochondria (IFM) are positioned between myofibrils that wrap around the I-band of the sarcomere (Bleck et al., 2018) and form extensive networks that facilitate rapid distribution of ATP to myofibrillar ATPases to support muscle contraction (Willingham et al., 2021). Peripherally located mitochondria (PM) localize to capillaries (paravascular mitochondria; PVM) or nuclei (paranuclear; PNM) (Glancy et al., 2014; Glancy et al., 2015) and are less branched and larger than IMF (Picard et al., 2013), which increases matrix and cristae volume. However, the bioenergetic differences between PVM and PNM remain unclear.
Electron transport chain
Four transmembrane protein complexes and two electron carriers are the basis for generating the mitochondrial membrane potential ( Figure 2 ). Through a series of redox reactions, electrons are transported from complex to complex while pumping protons from the matrix to the IMS. Complex I (CI) and complex II (CII) oxidize the electron carriers NADH or FADH2, respectively, to reduce the electron carrier ubiquinol. In the process, CI pumps protons across the IMM, whereas CII does not. Complex III (CIII) and complex IV (CIV) then catalyze the electron transfer from ubiquinol to cytochrome c and from cytochrome c to oxygen, respectively, to produce water and pump protons. Protons in the IMS can then re-enter the matrix through the F1F0 ATPase, also known as complex V (CV), where the energy generated from protons flowing down a concentration gradient is harnessed to phosphorylate ADP to produce ATP, termed OxPhos. The majority of proton re-entry is coupled to ATP production; however, alternative routes of re-entry exist that bypass CV, such as uncoupling proteins to produce heat or the nicotinamide nucleotide transhydrogenase to replenish the NADPH pool for biosynthetic processes ( Figure 2 ).
Although ETC complexes are routinely simplified as linear complexes in numerical order with a 1:1 ratio of complexes, their stoichiometry varies considerably (Schagger and Pfeiffer, 2001). Within the IMM folds, or cristae, complexes can be found as freely floating or in super-assembled structures known as supercomplexes. In mammalian cells, supercomplexes are largely composed of a combination of CI, CIII, and CIV, whereas the contribution of CII and CV to the formation of supercomplexes remains poorly understood (Acin-Perez et al., 2008). Their formation, although not fully understood, have been suggested to play a role in complex stability (Acin-Perez et al., 2004; Diaz et al., 2006), electron transport efficiency (Bianchi et al., 2004), and control of ROS production (Ghelli et al., 2013; Maranzana et al., 2013). Interestingly, supercomplex formation increases with exercise (Greggio et al., 2017), indicating that supercomplexes may be an adaptive mechanism to meet increased energy demands.
Mitochondrial dynamics
In skeletal muscle, mitochondria morphology and functions are quite responsive to a range of intermittent and continuous stimuli including exercise (Menshikova et al., 2006), nutritional factors (Mishra et al., 2015), aging (Coggan et al., 1992; Cooper et al., 1992; Short et al., 2005), and disease (Simoneau et al., 1995; Mootha et al., 2003; Patti et al., 2003). Mitochondrial adaptations include the ability to alter their abundance, ETC activity (oxidative capacity), coupling, and structure. Of those, the most rapid and fluid response is the change of structure (Eisner et al., 2014), which allows for dynamic mitochondrial remodeling to adapt to the cellular environment (Glancy et al., 2015). Mitochondrial dynamics are a coalition of fusion and fission events that are important for mitochondrial biogenesis, stress mitigation, and quality control by enabling the removal of damaged mitochondria (mitophagy; Figure 3 ). Fusion is mediated by mitofusion proteins (Mfn1 and Mfn2) and optic atrophy protein 1 (Opa1) to fuse the inner and outer mitochondrial membranes, respectively (Koshiba et al., 2004; Meeusen et al., 2004; Meeusen et al., 2006; Song et al., 2009). Deletion of Mfn1 and Mfn2 in skeletal muscle decreases oxidative capacity, reduces mitochondrial DNA (mtDNA) abundance, reduces muscle mass, and eventually leads to lethality (Chen et al., 2010a). On the other hand, fission events require the translocation of dynamin-related protein 1 (Drp1) from the cytosol to wrap around the outer membrane and cleave it (Ji et al., 2015; Kalia et al., 2018; Fonseca et al., 2019). Although there are other proteins involved in fission, loss of Drp1 inhibits fission and results in hyperfusion of mitochondria (Fonseca et al., 2019). Skeletal muscle specific ablation of Drp1 results in muscle atrophy, distorted mitochondrial morphology, and impaired mitochondrial respiration (Favaro et al., 2019). Although a physiological decrease in Drp1 may facilitate hypertrophy, as overloaded muscle had a decrease in Drp1, increase in fusion proteins, and greater mitochondrial area (Uemichi et al., 2021). Together, the balance of proper mitochondrial fusion and fission is essential for skeletal muscle homeostasis and may play a role in muscle hypertrophy.
An overview of mitochondrial dynamics. Skeletal muscle mitochondria form a dynamic network that is constantly adapting to changing external and internal stimuli. Punctate mitochondria can fuse together to form an interconnected network permitting an increase in oxidative capacity, dispersal of mitochondrial damage to promote survival, or an exchange of mitochondrial content (mtDNA and metabolites). Contrarily, mitochondria can undergo fission to remove damaged mitochondria through mitophagy, facilitate mitochondrial trafficking, or adapt to metabolic changes. Created with BioRender.com.
Mitochondrial dynamics also permit a fluid response to changing nutritional factors. For example, muscle cells isolated from the extensor digitorum longus (EDL) incubated in media containing acetoacetate for 24 h, an oxidative substrate, have more extensive (fused) mitochondrial networks than those treated with glucose for the same amount of time. Fused mitochondrial networks are predominant in oxidative fibers and have high rates of fusion, whereas a greater proportion of punctate mitochondria are found in glycolytic fibers (Mishra et al., 2015). Matrix content is shared in mitochondrial networks, which enables rapid cellular energy distribution (Glancy et al., 2015) and movement of metabolites, ions, proteins, and mtDNA. Thus, extensive mitochondrial networks in oxidative fibers facilitate energy homeostasis, whereas glycolytic fibers have less extensive networks that support short bouts of work through glycolysis. On the other hand, in conditions of nutrient deprivation, mitochondria elongate by downregulating fission events to protect against mitophagy and maximize energy production (Gomes et al., 2011). These findings demonstrate mitochondria readily adapt to support muscle physiology under a wide range of nutrient conditions.
Molecular regulators of skeletal muscle growth and metabolism
Skeletal muscle growth is orchestrated by a hierarchy of complex anabolic and catabolic biological processes that dictate the rate of protein accretion. Skeletal muscle is composed of a heterogenous population of muscle cells (myofibers) that have variable rates of myofibrillar, mitochondrial, and sarcoplasmic protein turnover. As such, protein accretion rate is a function of several myofiber characteristics. Myofibers can be classified according to predominate type of energy metabolism (glycolytic or oxidative metabolism) and speed of contraction (fast or slow). Slow-twitch muscle fibers have a greater rate of protein synthesis than fast-twitch fibers which can be attributed to duration of contraction and mitochondrial abundance, as mitochondrial proteins have higher fractional synthesis rates than myofibrillar proteins (Jaleel et al., 2008). However, slow fibers also have high rates of protein degradation, presumably due to a high mitochondrial content prone to damage from reactive oxygen species, that has a net result of low rates of protein accretion (Lewis et al., 1984) and relatively small fiber size. Therefore, metabolism is a contributing factor to muscle accretion rate and can be exploited to manipulate growth rate and efficiency of substrate utilization. Indeed, skeletal muscle from high-feed efficient pigs have a greater propensity of glycolytic metabolism than low feed efficient pigs (Fu et al., 2017). However, the regulatory mechanisms that facilitate metabolic shifts to promote an improved feed conversion ratio remain elusive.
IGF-Akt signaling
The two major regulatory pathways of protein accretion in muscle are the insulin-like growth factor 1 (IGF-1)/ phosphoinositide-3-kinase-Akt (PI3K-Akt)/ protein kinase B (PKB)/ mammalian target of rapamycin (mTOR) pathway and the myostatin pathway, which are respective positive and negative regulators of protein synthesis. Binding of IGF-1 to its receptor activates PI3K, which phosphorylates phosphoinositide-4,5-biphosphate (PIP2) generating phosphoinositide-3,4,5-triphosphate (PIP3) to act as a docking site for phosphoinositide-dependent kinase 1 (PDPK1) to activate Akt. Activation of Akt inhibits protein degradation through the repression of the forkhead box protein (FOXO) family of transcription factors and stimulates protein synthesis through the activation of mTOR and glycogen synthase kinase 3β (GSK3β). In skeletal muscle, loss of the IGF-1 receptor induces muscle atrophy (Mavalli et al., 2010), whereas IGF-1 overexpression stimulates hypertrophy (Coleman et al., 1995; Musaro et al., 2001; Fiorotto et al., 2003). Similarly, the activation of its downstream target Akt in skeletal muscle induces profound hypertrophy (Lai et al., 2004; Blaauw et al., 2009), whereas deletion of the downstream target mTOR results in impaired postnatal growth and the development of severe myopathy (Risson et al., 2009). Furthermore, loss of mTOR1 diminishes oxidative capacity of soleus and EDL muscles, which suggests that mTOR1 plays a role in modulating metabolic characteristics of skeletal muscle (Bentzinger et al., 2008). As diminished oxidative capacity is often associated with the onset of numerous metabolic diseases and muscle wasting, it is plausible this metabolic shift plays a role in the development of myopathies in muscle lacking mTOR1. Although it is apparent oxidative metabolism is key to longevity and protection against metabolic disease in humans, scientists in animal agriculture have a vastly different objective of facilitating efficient lean accretion rather than alleviating ailments associated with aging. However, these discoveries are still of importance to animal agriculture. For example, overexpression of IGF-1 results in heavier body weights, muscle mass, and a shift to carbohydrates as the primary fuel source (Christoffolete et al., 2015). However, oxygen consumption did not change (Christoffolete et al., 2015), which suggests that IGF-1 mediated hypertrophy is fueled by carbohydrate oxidation. Together, these findings suggest that IGF-1 signaling prioritizes glucose as a fuel source to facilitate hypertrophy, and although it seems that mitochondria play a role, a deeper understanding of this junction is necessary to translate these findings to develop innovative strategies for efficient muscle hypertrophy in meat producing animals.
Myostatin signaling
Myostatin is a negative regulator of muscle growth. Natural mutations of the myostatin gene have been discovered in sheep (Clop et al., 2006), cattle (Grobet et al., 1997; McPherron and Lee, 1997), dogs (Mosher et al., 2007), and humans (Schuelke et al., 2004) that result in gene disruption and a subsequent increase in muscle mass. In mice, myostatin deletion results in a 2-fold increase in muscle mass and decreased adipose tissue deposition (McPherron and Lee, 2002; Hamrick et al., 2006), whereas myostatin overexpression results in a reduction of muscle mass (Reisz-Porszasz et al., 2003). Myostatin deficiency also improves glucose metabolism and decreases mitochondrial respiration (Chen et al., 2010b; Mouisel et al., 2014). Myostatin and IGF-1 have a dynamic interplay that modulates muscle accretion through the two pathways. The PI3K/Akt axis is common to both pathways and permits crosstalk, which is essential for normal muscle growth and physiology. For example, overexpression of myostatin downregulates Akt/mTOR signaling and subsequently decreases protein synthesis (Amirouche et al., 2009). Contrarily, myostatin deletion increases PKB and mTOR activity (Lipina et al., 2010), although presence of the IGF-1 receptor is necessary for efficient hypertrophy (Kalista et al., 2012). Together, both pathways play a role of modulating metabolism and are pertinent in the regulation of muscle hypertrophy.
AMPK signaling
Although proper myostatin and IGF-1 crosstalk are essential for muscle physiology and growth, it is apparent the maintenance of metabolic homeostasis is also crucial. Regulation of skeletal muscle metabolism involves highly integrated signaling pathways that are tightly coupled to nutrient availability and cellular energy status. One example is the dynamic intracellular energy sensor AMPK that regulates muscle metabolism by relaying signals of energetic stress. As ATP is hydrolyzed, AMP levels rise and decrease the ATP:AMP ratio to activate AMPK, which promotes catabolic processes and suppresses anabolic processes to replenish ATP stores. As such, activation of AMPK decreases protein synthesis, an anabolic process, demonstrated by a 45% decrease in skeletal muscle fractional rate of protein synthesis 1 h after injection of the AMPK activating drug 5-aminoimidazole-4-carboxamide ribonucleotide (AICAR) (Bolster et al., 2002). AMPK activation decreases protein accretion rates by suppressing mTOR activity and activating FOXO and uncoordinated 51-like kinase 1 (ULK1), regulators of the autophagy pathway (Lee et al., 2010; Kim et al., 2011). Downregulation of anabolic pathways from AMPK activation is also accompanied by greater catabolism to generate ATP and meet the increased energy demand. For example, mutation of the AMPKγ3 subunit in mice (Garcia-Roves et al., 2008) and pigs (Scheffler et al., 2014) results in constitutive activation of AMPK, which increases mitochondrial biogenesis and enhances oxidative capacity of glycolytic muscle. Surprisingly, its chronic activation does not impact muscle growth (Carr et al., 2006), which may be attributed to a compensatory or adaptive mechanism that can override the negative regulation of AMPK on protein synthesis. In fact, Scheffler et al. (2014) found that the AMPKγ3 R200Q mutant pigs had an increase in Akt expression, a positive regulator of muscle growth that may account for the normal growth of these pigs. In agreement, Lantier et al. (2010) reported a skeletal muscle-specific double knockout of AMPKα1 and AMPKα2 increased soleus muscle mass, which further implicates AMPK as a regulator of muscle mass. As a result, the AMPK pathway has become a target to counteract metabolic disorders without negatively impacting muscle mass, as an increase in oxidative capacity enhances the ability of muscle to combat energetic stress providing protection against muscle wasting and metabolic diseases (Conley et al., 2000; Wang et al., 2004; Wenz et al., 2008; Wenz et al., 2009; Ljubicic et al., 2011). Indeed, an increase in oxidative capacity can occur simultaneously with hypertrophy (Scheffler et al., 2014), which suggests that a compensatory pathway exists that can force hypertrophy regardless of metabolic characteristics that has yet to be discovered. In summary, signaling pathways that modulate muscle hypertrophy are complex and intricately interlaced with those that regulate skeletal muscle metabolism, which has downstream implications on lean accretion efficiency.
Nutrient sensing and skeletal muscle adaptation
O-GlcNAcylation
Muscle cells can perceive cellular energy status and substrate availability, and as such, muscle is a plastic tissue that permits metabolic adaptations to environmental changes. During times of nutrient surplus, such as those after feeding, muscle can “sense” the availability of substrates and increase protein accretion. Conversely, periods of nutrient scarcity result in repartition of nutrients away from anabolic processes. The dynamic post-translational modification O-linked-β-D-N-acetylglucosamine (O-GlcNAc) serves as a widespread nutrient gauge by cycling on and off O-GlcNAc moieties to serine and threonine residues of nuclear, cytosolic, and mitochondrial proteins (Hanover et al., 2012). The donor substrate, UDP-GlcNAc, is a product of the HBP and its levels reflect the collective cellular nutrient flux by incorporating nucleotide, carbohydrate, lipid, and protein metabolism. A pair of enzymes, O-GlcNAc transferase (OGT) and O-GlcNAcase (OGA), are responsible for the respective addition and removal of O-GlcNAc from target proteins to modify their function to adapt to cellular nutrient abundance ( Figure 4 ). Collectively, this post-translational modification acts as a nutrient sensing pathway because levels of UDP-GlcNAc, and thus O-GlcNAcylation, are indicative of cellular nutrient status (Walgren et al., 2003; Housley et al., 2008; Kang et al., 2008).
O-GlcNAc signaling pathway. Step 1: the hexosamine biosynthetic pathway incorporates metabolites from protein, carbohydrate, lipid, energy, and nucleotide metabolism to synthesize UDP-GlcNAc, which fluctuates with cellular nutrient status. Step 2: OGT transfers UDP-GlcNAc to any of its thousands of target proteins and OGA removes the O-GlcNAc moiety, known as O-GlcNAcylation, to modulate protein function. O-GlcNAcylation occurs in a similar manner to phosphorylation and may be as widespread, but O-GlcNAcylation patterns correlate to cellular nutrient status and is only mediated through the enzymes OGT and OGA. In skeletal muscle, glycolytic muscle has a greater abundance of O-GlcNAcylated proteins. In fact, several glycolytic enzymes and mitochondrial proteins are targets of OGT, and as such O-GlcNAcylation can modulate skeletal muscle metabolism in response to changing dietary conditions. Created with BioRender.com.
O-GlcNAc transferase
A single gene encodes OGT; however, three isoforms originate from alternative splicing. The two predominant isoforms are nucleocytoplasmic OGT (ncOGT) and mitochondrial OGT (mOGT), which contains an N-terminal mitochondrion targeting sequence. A short isoform of OGT exists, but its functionality remains largely unclear. Until recently, global cellular O-GlcNAcylation has been largely accredited to ncOGT; however, Banerjee et al. (2015) showed that the pyrimidine nucleotide carrier transports UDP-GlcNAc into the matrix and blocking this transporter lowers mitochondrial O-GlcNAcylation, which suggests that robust O-GlcNAc cycling occurs within mitochondria. Sacoman and colleagues (2017) were able to knock down cellular OGT and mOGT in HeLa cells by targeting unique sequences of the splice variants. Knockdown of mOGT altered mitochondrial content, function, and membrane potential; however, knockdown of both mOGT and ncOGT increased mitochondrial content and restored membrane potential. These data suggest that crosstalk between the two isoforms of OGT exists to mediate mitochondrial integrity and function, yet its relation to nutrient status remains elusive.
OGT substrates
Several metabolic enzymes are targets of O-GlcNAcylation. Phosphofructokinase-1, a rate-limiting glycolytic enzyme, is inhibited by O-GlcNAcylation resulting in the redirection of glucose to the PPP in cancer cells (Yi et al., 2012). O-GlcNAcylation of G6PDH, the rate limiting enzyme of the PPP, increases its activity and glucose flux through the PPP in lung cancer cells (Rao et al., 2015). Prolonged exposure of cells to glucose enhances cellular O-GlcNAcylation and subsequently increases glycolytic flux (Marshall et al., 1991; Hanover et al., 1999; Bond and Hanover, 2015). In cardiomyocytes, Hu et al. (2009) reported several subunits of the electron transport chain are O-GlcNAcylated in high glucose conditions, which decrease complex activity and mitochondrial respiration (Hu et al., 2009). Interestingly, removal of O-GlcNAcylation in high glucose conditions restores mitochondrial function to that of cells exposed to normal glucose. These data show that elevated glucose availability signals an increase in O-GlcNAcylation of key regulatory protens to modulate metabolism.
HBP pathway
Glucose and glutamine are both considered rate limiting in the HBP, and as such are potent stimulators of UDP-GlcNAc synthesis and thus O-GlcNAcylation (Walgren et al., 2003; Swamy et al., 2016). Excess free fatty acids also increase UDP-GlcNAc levels in rat skeletal muscle, although the subsequent glucose uptake and handling are markedly different than that stimulated by excess glucose (Hawkins et al., 1997b). In cultured pancreatic beta-cells, high-fat exposure increased UDP-GlcNAc levels and resulted in dysregulated metabolites distinct from those exposed to high glucose (Yousf et al., 2019), which demonstrates that UDP-GlcNAc levels increase in response to different nutrients but their downstream metabolic adaptations vary uniquely.
O-GlcNAcylation and muscle metabolism
O-GlcNAcylation has been well documented in skeletal muscle and has emerged to have an important role in maintaining proper skeletal muscle glucose handling and insulin sensitivity. As skeletal muscle is the largest metabolically active tissue in the body and is responsible for the majority of insulin-mediated glucose disposal, skeletal muscle O-GlcNAcylation is a key player in global metabolism. Early investigation in muscle suggested a correlation between nutrient abundance and insulin resistance in rodents. Indeed, overexpression of GLUT4 in skeletal muscle results in chronic exposure of muscle to glucose, which is accompanied by greater UDP-GlcNAc levels and the development of insulin resistance in mice (Buse et al., 1996). Furthermore, prolonged fatty acid and uridine exposure increase UDP-GlcNAc levels and reduce skeletal muscle insulin sensitivity (Hawkins et al., 1997a; Hawkins et al., 1997b). Contrarily, ablation of OGT, and thus O-GlcNAcylation, in skeletal muscle protects mice from high fat diet induced obesity and insulin resistance (Shi et al., 2018). These findings demonstrate that O-GlcNAcylation in muscle has implications on whole-body metabolism.
Implications of skeletal muscle O-GlcNAcylation in farm animals
In cattle, O-GlcNAcylation is greater in animals on a grain-based diet than those on a forage-based diet (Apaoblaza et al., 2020). As O-GlcNAcylation is an indicator of high nutrient availability, the post-translation modification is presumably a key player in the metabolic shift towards glycolytic metabolism in cattle fed a concentrate diet. Contrarily, forage-based diets that are energy scarce promote oxidative metabolism and have low levels of O-GlcNAcylation. In pigs, low- and high- energy diets also result in distinct patterns of O-GlcNAcylated proteins where high-energy diets have high levels of O-GlcNAcylation (data not published). Collectively, these data document the existence of intact hexamine biosynthesis signaling and O-GlcNAcylation in muscle of meat producing animals and strongly argue that O-GlcNAcylation is part of the complex nutrient sensing machinery that modulates both skeletal muscle and whole body metabolism.
Metabolic and nutritional regulation of satellite cells
Adult skeletal muscle contains a heterogeneous population of stem cells, known as satellite cells (SC), that contribute to postnatal muscle growth, maintenance, and repair (Mauro, 1961; Moss and Leblond, 1971; Kuang et al., 2007; Lepper et al., 2011). Adult SC are predominately in a quiescent state and can activate to support several rounds of proliferation before undergoing self-renewal or terminal differentiation (Schultz et al., 1978; Snow, 1978). Orchestration of these behaviors is tightly synchronized by integrated signaling cascades that relay unfolding intrinsic and extrinsic stimuli.
Metabolic regulation of satellite cells
Intrinsically, metabolism is a strong regulator of SC behavior, during which metabolic shifts occur to accommodate demands of different growth stages ( Figure 5 ). Quiescent SC have a relatively low metabolic rate but can be separated into two distinct metabolic populations. The population with a lower metabolic rate has a delayed cell cycle entry, a slower mitotic rate, and are less transcriptionally primed for myogenic commitment. Contrarily, the population with higher metabolic activity is more responsive to activation (Rocheteau et al., 2012). Once activated, SC begin to proliferate, accompanied by an increase in glycolysis and glutaminolysis to meet energy demands and support macromolecule biosynthesis, which is essential for proliferating cells (Lunt and Vander Heiden, 2011). Within 48 h of differentiation, mitochondrial content and oxidative capacity increase to shift to a more oxidative metabolism (Moyes et al., 1997; Remels et al., 2010). Disruption of this shift through inhibition of mitochondrial protein synthesis (Korohoda et al., 1993; Hamai et al., 1997; Rochard et al., 2000), mitochondrial fission (Bloemberg and Quadrilatero, 2016), mitochondrial DNA replication (Brunk and Yaffe, 1976; Herzberg et al., 1993), or inhibition of OxPhos confines myoblast to a proliferative, single cell state and obstructs myoblast differentiation. Although an oxidative shift is necessary to support differentiation, glucose is still an essential nutrient, which suggests glucose oxidation supports myogenesis. Indeed, obstruction of glucose entry into the TCA cycle through deletion of PDH impairs differentiation (Hori et al., 2019). These findings define the metabolic changes that facilitate myogenesis.
Metabolic changes associated with adult myogenesis. Quiescent SC have minimal metabolic activity and few mitochondria. As SC activate, a rapid increase in metabolic activity takes place that is accompanied by an increase in mitochondrial content, which continues to increase as SC exit the cell cycle to differentiate. In fact, obstruction of mitochondrial biogenesis impedes myogenesis. Images are from isolated and cultured single myofibers that were fixed and incubated in Mitotracker Red (red) and an antibody against Pax7 (green) then counterstained with DAPI (blue) created with BioRender.com.
AMPK signaling
Changes in cellular energy status also have implications on myogenesis. AMPK, activated by changes in AMP:ATP ratios, is a sensor of metabolic stress that signals periods of energy stress to facilitate adaptations to changing internal conditions. Activation of AMPK with the drug AICAR impairs G1/S cell cycle progression and reduces myotube formation (Williamson et al., 2007). On the other hand, deletion of AMPKα1 in SC impairs adult regenerative myogenesis, and SC lacking AMPKα1 transplanted into control muscles have a diminished myogenic capacity (Fu et al., 2016), which suggests that external conditions cannot rescue the loss of AMPKα1 in SC. In addition to impaired regeneration, these cells proliferate more slowly in culture (Fu et al., 2015), which can be attributed to the loss of AMPK obstructing Warburg-like glycolysis. In proliferating cells, an increase in Warburg-like glycolysis is stimulated by AMPK through a decrease in ATP to ADP ratios to support proliferation, whereas non-proliferating cells have a higher ratio of ATP to ADP accompanied by greater mitochondrial function (Maldonado and Lemasters, 2014). Collectively, these findings highlight the importance of AMPK translating cellular energy status to modulate myoblast metabolism for efficient myogenic progression.
Nutrient conditions on satellite cell fate
Extrinsically, conditions of nutrient excess or scarcity are also perceived by SC and alter their behavior accordingly. In SC, O-GlcNAcylation, the widespread nutrient sensing pathway, is essential for normal SC behavior. Loss of OGT, and thus O-GlcNAcylation, impairs SC proliferation and adult regenerative myogenesis (Zumbaugh et al., 2021). More specifically, high circulating levels of amino acids promote SC lineage progression through the mTOR pathway (Dai et al., 2015), a cellular nutrient sensor that regulates protein synthesis rates. For example, supplementation of lysine, the first limiting amino acid in high carbohydrate diets (Li et al., 2012; Zeng et al., 2013), increases mTOR activity, which suppress proteolysis through the autophagic-lysosomal system (Sato et al., 2014) and increases SC proliferation (Jin et al., 2019). Furthermore, leucine supplementation promotes proliferation, differentiation, and skeletal muscle regeneration through the mTORC1 pathway (Pereira et al., 2014; Dai et al., 2015), and leucine restriction inhibitions differentiation (Averous et al., 2012). Indeed, inhibition of mTOR in myoblasts cultured in high leucine concentrations decreases proliferative capacity and protein synthesis (Han et al., 2008). In agreement, conditional deletion of mTOR in adult SC also impairs proliferation, differentiation, and muscle regeneration (Zhang et al., 2015). Collectively, these findings show that the mTOR pathway is essential to translate fluctuations in nutrient abundance to modulate myoblast behavior. In addition to promoting protein synthesis to support myogenesis, mTOR also regulates myogenesis through the expression of several myogenic regulatory factors including Myf5 and MyoD (Averous et al., 2012; Hatfield et al., 2015; Zhang et al., 2015), which play a role in SC activation and transient proliferation (Braun et al., 1992; Rudnicki et al., 1992; Ustanina et al., 2007), and myogenin, which is required for terminal myoblast differentiation (Hasty et al., 1993; Nabeshima et al., 1993). Collectively, these findings demonstrate the importance of adequate nutrient availability and the downstream signaling that indicates cellular nutrient status for proper SC function.
On the other hand, overfeeding or excessive substrate availability negatively impacts SC function (Peterson et al., 2008; D’Souza et al., 2015; Fausnacht et al., 2020). For example, high levels of glucose induce insulin resistance in C2C12 myotubes and inhibit myogenic differentiation (Grzelkowska-Kowalczyk et al., 2013), which is accompanied by a decrease in Akt phosphorylation (Luo et al., 2019). In fact, inhibition of Akt (Luo et al., 2019) or PI3-kinase (Tureckova et al., 2001) recapitulates the high glucose phenotype and inhibits differentiation. As such, activation of the PI3-kinase/Akt pathway is essential for efficient glucose uptake. Indeed, pharmacological activation of Akt in high glucose conditions rescues myogenic differentiation (Luo et al., 2019). These findings demonstrate the PI3-kinase/Akt pathway can “sense” changes in nutrient concentrations to alter SC behavior. Alternatively, short-term calorie restriction improves SC function and increases oxidative metabolism (Cerletti et al., 2012). However, discrepancies in the effectiveness and response to caloric restriction have been debated in regard to sex, age, and duration of restriction (Boldrin et al., 2017; Abreu et al., 2020). In summary, SC are susceptible to changing extrinsic nutrient and intrinsic metabolic conditions through complex signaling pathways that modulate myogenesis.
Summary
Skeletal muscle is a heterogenous population of myofibers and SC with varying metabolic characteristics that influence muscle physiology and myogenesis, respectively. Skeletal muscle metabolism is modulated by several signaling pathways that have implications on fiber size, nutrient utilization, and protein turnover rate. Additionally, myogenesis is governed by distinct metabolic adaptations and signaling pathways that are responsive to changing external conditions. Further investigation into the influence of skeletal muscle and SC metabolic regulation on livestock growth efficiency will provide the framework for an animal scientist to develop selection strategies that may prove useful in augmenting animal growth efficiency and meat production.
Acknowledgments
All figures made using BioRender.com. An acknowledgement for USDA Proposal Number: 2018-07081
Glossary
Abbreviations
ADP | adenosine diphosphate |
AICAR | 5-aminoimidazole-4-carboxamide ribonucleotide |
Akt1 | AKT serine/threonine kinase 1 |
AMP | adenosine monophosphate |
AMPK | 5ʹ-adenosine monophosphate-activated protein kinase |
ATP | adenosine triphosphate |
Cr | creatine |
CI | complex I |
CII | complex II |
CIII | complex III |
CIV | complex IV |
CV | complex V |
Drp1 | dynamin-related protein 1 |
EDL | extensor digitorum longus |
ETC | electron transport chain |
FADH | flavin adenine dinucleotide |
FE | feed efficiency |
FOXO | forkhead box protein |
F6P | fructose-6-phosphate |
GFAT | fructose 6-phosphate amidotransferase |
GLUT4 | glucose transporter type 4 |
GSK3β | glycogen synthase kinase 3β |
G6P | glucose-6-phosphate |
G6PDH | glucose-6-phosphate dehydrogenase |
HBP | hexosamine biosynthetic pathway |
IGF-1 | insulin-like growth factor 1 |
IMM | inner mitochondrial membrane |
IMS | inner membrane space |
LDH | lactate dehydrogenase |
Mfn1 | mitofusion protein 1 |
Mfn2 | mitofusion protein 2 |
mOGT | mitochondrial O-GlcNAc transferase |
mtDNA | mitochondrial DNA |
mTOR | mammalian target of rapamycin |
NADH | nicotinamide adenine dinucleotide |
NADPH | nicotinamide adenine dinucleotide phosphate |
ncOGT | nucleocytoplasmic O-GlcNAc transferase |
O-GlcNAc | O-linked-β-D-N-acetylglucosamine |
O-GlcNAcylation | O-linked-N-acetylglucosaminylation |
OGA | O-GlcNAcase |
OGT | O-GlcNAc transferase |
OMM | outer mitochondrial membrane |
Opa1 | optic atrophy protein 1 |
OxPhos | oxidative phosphorylation |
PCr | phosphocreatine |
PDC | pyruvate dehydrogenase complex |
PDH | pyruvate dehydrogenase |
PDK | pyruvate dehydrogenase kinase |
PDK4 | pyruvate dehydrogenase kinase 4 |
PDP | pyruvate dehydrogenase phosphatases |
PDPK1 | phosphoinositide-dependent kinase 1 |
PFK | phosphofructokinase |
PGC-1α | peroxisome proliferator-activated receptor gamma coactivator 1-alpha |
PI3K | phosphoinositide-3-kinase |
PIP2 | phosphoinositide-4,5-biphosphate |
PIP3 | phosphoinositide-3,4,5-triphosphate |
PKB | protein kinase B |
PPAR | peroxisome proliferator-activated receptors |
PPP | pentose phosphate pathway |
ROS | reactive oxygen species |
SC | satellite cells |
SR | sarcoplasmic reticulum |
TCA | tricarboxylic acid |
UDP-GlcNAc | uridine diphosphate N-acetylglucosamine |
ULK1 | uncoordinated 51-like kinase 1 |
6PGGDH | 6-phosphogluconate dehydrogenase |
Contributor Information
Morgan D Zumbaugh, Department of Animal and Poultry Sciences, Virginia Polytechnic Institute and State University, Blacksburg, VA 24061, USA.
Sally E Johnson, Department of Animal and Poultry Sciences, Virginia Polytechnic Institute and State University, Blacksburg, VA 24061, USA.
Tim H Shi, Department of Animal and Poultry Sciences, Virginia Polytechnic Institute and State University, Blacksburg, VA 24061, USA.
David E Gerrard, Department of Animal and Poultry Sciences, Virginia Polytechnic Institute and State University, Blacksburg, VA 24061, USA.
Conflict of Interest Statement
The authors declare no real or perceived conflicts of interest.
Literature Cited
- Abreu, P., Serna J. D. C., Munhoz A. C., and Kowaltowski A. J.. . 2020. Calorie restriction changes muscle satellite cell proliferation in a manner independent of metabolic modulation . Mech. Ageing Dev . 192 :111362. doi: 10.1016/j.mad.2020.111362. [PubMed] [CrossRef] [Google Scholar]
- Acín-Pérez, R., Bayona-Bafaluy M. P., Fernández-Silva P., Moreno-Loshuertos R., Pérez-Martos A., Bruno C., Moraes C. T., and Enríquez J. A.. . 2004. Respiratory complex III is required to maintain complex I in mammalian mitochondria . Mol. Cell 13 :805–815. doi: 10.1016/s1097-2765(04)00124-8. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Acín-Pérez, R., Fernández-Silva P., Peleato M. L., Pérez-Martos A., and Enriquez J. A.. . 2008. Respiratory active mitochondrial supercomplexes . Mol. Cell 32 :529–539. doi: 10.1016/j.molcel.2008.10.021. [PubMed] [CrossRef] [Google Scholar]
- Ahn, C. S., and Metallo C. M.. . 2015. Mitochondria as biosynthetic factories for cancer proliferation . Cancer Metab . 3 :1. doi: 10.1186/s40170-015-0128-2. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Amirouche, A., Durieux A. C., Banzet S., Koulmann N., Bonnefoy R., Mouret C., Bigard X., Peinnequin A., and Freyssenet D.. . 2009. Down-regulation of Akt/mammalian target of rapamycin signaling pathway in response to myostatin overexpression in skeletal muscle . Endocrinology 150 :286–294. doi: 10.1210/en.2008-0959. [PubMed] [CrossRef] [Google Scholar]
- Aon, M. A., Cortassa S., and O’Rourke B.. . 2010. Redox-optimized ROS balance: a unifying hypothesis . Biochim. Biophys. Acta 1797 :865–877. doi: 10.1016/j.bbabio.2010.02.016. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Apaoblaza, A., Gerrard S. D., Matarneh S. K., Wicks J. C., Kirkpatrick L., England E. M., Scheffler T. L., Duckett S. K., Shi H., Silva S. L., . et al. 2020. Muscle from grass- and grain-fed cattle differs energetically . Meat Sci . 161 :107996. doi: 10.1016/j.meatsci.2019.107996. [PubMed] [CrossRef] [Google Scholar]
- Arnold, D. L., Matthews P. M., and Radda G. K.. . 1984. Metabolic recovery after exercise and the assessment of mitochondrial function in vivo in human skeletal muscle by means of 31P NMR . Magn. Reson. Med . 1 :307–315. doi: 10.1002/mrm.1910010303. [PubMed] [CrossRef] [Google Scholar]
- Averous, J., Gabillard J. C., Seiliez I., and Dardevet D.. . 2012. Leucine limitation regulates myf5 and myoD expression and inhibits myoblast differentiation . Exp. Cell Res . 318 :217–227. doi: 10.1016/j.yexcr.2011.10.015. [PubMed] [CrossRef] [Google Scholar]
- Banerjee, P. S., Ma J., and Hart G. W.. . 2015. Diabetes-associated dysregulation of O-GlcNAcylation in rat cardiac mitochondria . Proc. Natl. Acad. Sci. U.S.A . 112 :6050–6055. doi: 10.1073/pnas.1424017112. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Battistuzzi, G., D’Urso M., Toniolo D., Persico G. M., and Luzzatto L.. . 1985. Tissue-specific levels of human glucose-6-phosphate dehydrogenase correlate with methylation of specific sites at the 3’ end of the gene . Proc. Natl. Acad. Sci. U.S.A . 82 :1465–1469. doi: 10.1073/pnas.82.5.1465. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Bentzinger, C. F., Romanino K., Cloëtta D., Lin S., Mascarenhas J. B., Oliveri F., Xia J., Casanova E., Costa C. F., Brink M., . et al. 2008. Skeletal muscle-specific ablation of raptor, but not of rictor, causes metabolic changes and results in muscle dystrophy . Cell Metab . 8 :411–424. doi: 10.1016/j.cmet.2008.10.002. [PubMed] [CrossRef] [Google Scholar]
- Bianchi, C., Genova M. L., Parenti Castelli G., and Lenaz G.. . 2004. The mitochondrial respiratory chain is partially organized in a supercomplex assembly: kinetic evidence using flux control analysis . J. Biol. Chem . 279 :36562–36569. doi: 10.1074/jbc.M405135200. [PubMed] [CrossRef] [Google Scholar]
- Blaauw, B., Canato M., Agatea L., Toniolo L., Mammucari C., Masiero E., Abraham R., Sandri M., Schiaffino S., and Reggiani C.. . 2009. Inducible activation of Akt increases skeletal muscle mass and force without satellite cell activation . FASEB J . 23 :3896–3905. doi: 10.1096/fj.09-131870. [PubMed] [CrossRef] [Google Scholar]
- Bleck, C. K. E., Kim Y., Willingham T. B., and Glancy B.. . 2018. Subcellular connectomic analyses of energy networks in striated muscle . Nat. Commun . 9 :5111. doi: 10.1038/s41467-018-07676-y. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Bloemberg, D., and Quadrilatero J.. . 2016. Effect of mitochondrial fission inhibition on C2C12 differentiation . Data Brief 7 :634–640. doi: 10.1016/j.dib.2016.02.070. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Bohovych, I., and Khalimonchuk O.. . 2016. Sending out an SOS: mitochondria as a signaling hub . Front. Cell Dev. Biol . 4 :109. doi: 10.3389/fcell.2016.00109. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Boldrin, L., Ross J. A., Whitmore C., Doreste B., Beaver C., Eddaoudi A., Pearce D. J., and Morgan J. E.. . 2017. The effect of calorie restriction on mouse skeletal muscle is sex, strain and time-dependent . Sci. Rep . 7 :5160. doi: 10.1038/s41598-017-04896-y. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Bolster, D. R., Crozier S. J., Kimball S. R., and Jefferson L. S.. . 2002. AMP-activated protein kinase suppresses protein synthesis in rat skeletal muscle through down-regulated mammalian target of rapamycin (mTOR) signaling . J. Biol. Chem . 277 :23977–23980. doi: 10.1074/jbc.C200171200. [PubMed] [CrossRef] [Google Scholar]
- Boncompagni, S., Rossi A. E., Micaroni M., Beznoussenko G. V., Polishchuk R. S., Dirksen R. T., and Protasi F.. . 2009. Mitochondria are linked to calcium stores in striated muscle by developmentally regulated tethering structures . Mol. Biol. Cell 20 :1058–1067. doi: 10.1091/mbc.e08-07-0783. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Bond, M. R., and Hanover J. A.. . 2015. A little sugar goes a long way: the cell biology of O-GlcNAc . J. Cell Biol . 208 :869–880. doi: 10.1083/jcb.201501101. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Bowker-Kinley, M. M., Davis W. I., Wu P., Harris R. A., and Popov K. M.. . 1998. Evidence for existence of tissue-specific regulation of the mammalian pyruvate dehydrogenase complex . Biochem. J . 329 (Pt 1) :191–196. doi: 10.1042/bj3290191. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Brand, M. D. 2010. The sites and topology of mitochondrial superoxide production . Exp. Gerontol . 45 :466–472. doi: 10.1016/j.exger.2010.01.003. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Braun, T., Rudnicki M. A., Arnold H. H., and Jaenisch R.. . 1992. Targeted inactivation of the muscle regulatory gene Myf-5 results in abnormal rib development and perinatal death . Cell 71 :369–382. doi: 10.1016/0092-8674(92)90507-9. [PubMed] [CrossRef] [Google Scholar]
- Bricker, D. K., Taylor E. B., Schell J. C., Orsak T., Boutron A., Chen Y. C., Cox J. E., Cardon C. M., Van Vranken J. G., Dephoure N., . et al. 2012. A mitochondrial pyruvate carrier required for pyruvate uptake in yeast, Drosophila, and humans . Science 337 :96–100. doi: 10.1126/science.1218099. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Brooks, G. A. 1986. The lactate shuttle during exercise and recovery . Med. Sci. Sports Exerc . 18 :360–368. doi: 10.1249/00005768-198606000-00019. [PubMed] [CrossRef] [Google Scholar]
- Brunk, C. F., and Yaffe D.. . 1976. The reversible inhibition of myoblast fusion by ethidium bromide (EB) . Exp. Cell Res . 99 :310–318. doi: 10.1016/0014-4827(76)90588-7. [PubMed] [CrossRef] [Google Scholar]
- Bulteau, A. L., O’Neill H. A., Kennedy M. C., Ikeda-Saito M., Isaya G., and Szweda L. I.. . 2004. Frataxin acts as an iron chaperone protein to modulate mitochondrial aconitase activity . Science 305 :242–245. doi: 10.1126/science.1098991. [PubMed] [CrossRef] [Google Scholar]
- Buse, M. G., Robinson K. A., Marshall B. A., and Mueckler M.. . 1996. Differential effects of GLUT1 or GLUT4 overexpression on hexosamine biosynthesis by muscles of transgenic mice . J. Biol. Chem . 271 :23197–23202. doi: 10.1074/jbc.271.38.23197. [PubMed] [CrossRef] [Google Scholar]
- Carr, C. C., Morgan J. B., Berg E. P., Carter S. D., and Ray F. K.. . 2006. Growth performance, carcass composition, quality, and enhancement treatment of fresh pork identified through deoxyribonucleic acid marker-assisted selection for the Rendement Napole gene . J. Anim. Sci . 84 :910–917. doi: 10.2527/2006.844910x. [PubMed] [CrossRef] [Google Scholar]
- Cerletti, M., Jang Y. C., Finley L. W., Haigis M. C., and Wagers A. J.. . 2012. Short-term calorie restriction enhances skeletal muscle stem cell function . Cell Stem Cell 10 :515–519. doi: 10.1016/j.stem.2012.04.002. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Chen, H., Vermulst M., Wang Y. E., Chomyn A., Prolla T. A., McCaffery J. M., and Chan D. C.. . 2010. Mitochondrial fusion is required for mtDNA stability in skeletal muscle and tolerance of mtDNA mutations . Cell 141 :280–289. doi: 10.1016/j.cell.2010.02.026. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Chen, Y., Ye J., Cao L., Zhang Y., Xia W., and Zhu D.. . 2010. Myostatin regulates glucose metabolism via the AMP-activated protein kinase pathway in skeletal muscle cells . Int. J. Biochem. Cell Biol . 42 :2072–2081. doi: 10.1016/j.biocel.2010.09.017. [PubMed] [CrossRef] [Google Scholar]
- Christoffolete, M. A., Silva W. J., Ramos G. V., Bento M. R., Costa M. O., Ribeiro M. O., Okamoto M. M., Lohmann T. H., Machado U. F., Musarò A., . et al. 2015. Muscle IGF-1-induced skeletal muscle hypertrophy evokes higher insulin sensitivity and carbohydrate use as preferential energy substrate . Biomed Res. Int . 2015 :282984. doi: 10.1155/2015/282984. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Clop, A., Marcq F., Takeda H., Pirottin D., Tordoir X., Bibé B., Bouix J., Caiment F., Elsen J. M., Eychenne F., . et al. 2006. A mutation creating a potential illegitimate microRNA target site in the myostatin gene affects muscularity in sheep . Nat. Genet . 38 :813–818. doi: 10.1038/ng1810. [PubMed] [CrossRef] [Google Scholar]
- Coggan, A. R., Spina R. J., King D. S., Rogers M. A., Brown M., Nemeth P. M., and Holloszy J. O.. . 1992. Histochemical and enzymatic comparison of the gastrocnemius muscle of young and elderly men and women . J. Gerontol . 47 :B71–B76. doi: 10.1093/geronj/47.3.b71. [PubMed] [CrossRef] [Google Scholar]
- Coleman, M. E., DeMayo F., Yin K. C., Lee H. M., Geske R., Montgomery C., and Schwartz R. J.. . 1995. Myogenic vector expression of insulin-like growth factor I stimulates muscle cell differentiation and myofiber hypertrophy in transgenic mice . J. Biol. Chem . 270 :12109–12116. doi: 10.1074/jbc.270.20.12109. [PubMed] [CrossRef] [Google Scholar]
- Conley, K. E., Jubrias S. A., and Esselman P. C.. . 2000. Oxidative capacity and ageing in human muscle . J. Physiol . 526 Pt 1 :203–210. doi: 10.1111/j.1469-7793.2000.t01-1-00203.x. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Cooper, J. M., Mann V. M., and Schapira A. H.. . 1992. Analyses of mitochondrial respiratory chain function and mitochondrial DNA deletion in human skeletal muscle: effect of ageing . J. Neurol. Sci . 113 :91–98. doi: 10.1016/0022-510x(92)90270-u. [PubMed] [CrossRef] [Google Scholar]
- D’Souza, D. M., Trajcevski K. E., Al-Sajee D., Wang D. C., Thomas M., Anderson J. E., and Hawke T. J.. . 2015. Diet-induced obesity impairs muscle satellite cell activation and muscle repair through alterations in hepatocyte growth factor signaling. Physiol. Rep . 3 ( 8 ). doi: 10.14814/phy2.12506 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Dai, J. M., Yu M. X., Shen Z. Y., Guo C. Y., Zhuang S. Q., and Qiu X. S.. . 2015. Leucine promotes proliferation and differentiation of primary preterm rat satellite cells in part through mTORC1 signaling pathway . Nutrients 7 :3387–3400. doi: 10.3390/nu7053387. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- DeFronzo, R. A., Jacot E., Jequier E., Maeder E., Wahren J., and Felber J. P.. . 1981. The effect of insulin on the disposal of intravenous glucose. Results from indirect calorimetry and hepatic and femoral venous catheterization . Diabetes 30 :1000–1007. doi: 10.2337/diab.30.12.1000. [PubMed] [CrossRef] [Google Scholar]
- Deluca, H. F., and Engstrom G. W.. . 1961. Calcium uptake by rat kidney mitochondria . Proc. Natl. Acad. Sci. U.S.A . 47 :1744–1750. doi: 10.1073/pnas.47.11.1744. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Depreux, F. F., Grant A. L., Anderson D. B., and Gerrard D. E.. . 2002. Paylean alters myosin heavy chain isoform content in pig muscle . J. Anim. Sci . 80 :1888–1894. doi: 10.2527/2002.8071888x. [PubMed] [CrossRef] [Google Scholar]
- Di Donato, D. M., West D. W., Churchward-Venne T. A., Breen L., Baker S. K., and Phillips S. M.. . 2014. Influence of aerobic exercise intensity on myofibrillar and mitochondrial protein synthesis in young men during early and late postexercise recovery . Am. J. Physiol. Endocrinol. Metab . 306 :E1025–E1032. doi: 10.1152/ajpendo.00487.2013. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Diaz, F., Fukui H., Garcia S., and Moraes C. T.. . 2006. Cytochrome c oxidase is required for the assembly/stability of respiratory complex I in mouse fibroblasts . Mol. Cell. Biol . 26 :4872–4881. doi: 10.1128/MCB.01767-05. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Donges, C. E., Burd N. A., Duffield R., Smith G. C., West D. W., Short M. J., Mackenzie R., Plank L. D., Shepherd P. R., Phillips S. M., . et al. 2012. Concurrent resistance and aerobic exercise stimulates both myofibrillar and mitochondrial protein synthesis in sedentary middle-aged men . J. Appl. Physiol. (1985) . 112 :1992–2001. doi: 10.1152/japplphysiol.00166.2012. [PubMed] [CrossRef] [Google Scholar]
- Eaton, S. 2002. Control of mitochondrial beta-oxidation flux . Prog. Lipid Res . 41 :197–239. doi: 10.1016/s0163-7827(01)00024-8. [PubMed] [CrossRef] [Google Scholar]
- Eisner, V., Lenaers G., and Hajnóczky G.. . 2014. Mitochondrial fusion is frequent in skeletal muscle and supports excitation-contraction coupling . J. Cell Biol . 205 :179–195. doi: 10.1083/jcb.201312066. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Fausnacht, D. W., McMillan R. P., Boutagy N. E., Lupi R. A., Harvey M. M., Davy B. M., Davy K. P., Rhoads R. P., and Hulver M. W.. . 2020. Overfeeding and substrate availability, but not age or BMI, alter human aatellite cell function. Nutrients 12 ( 8 ). doi: 10.3390/nu12082215 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Favaro, G., Romanello V., Varanita T., Andrea Desbats M., Morbidoni V., Tezze C., Albiero M., Canato M., Gherardi G., De Stefani D., . et al. 2019. DRP1-mediated mitochondrial shape controls calcium homeostasis and muscle mass . Nat. Commun . 10 :2576. doi: 10.1038/s41467-019-10226-9. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Fiorotto, M. L., Schwartz R. J., and Delaughter M. C.. . 2003. Persistent IGF-I overexpression in skeletal muscle transiently enhances DNA accretion and growth . FASEB J . 17 :59–60. doi: 10.1096/fj.02-0289fje. [PubMed] [CrossRef] [Google Scholar]
- Fonseca, T. B., Sánchez-Guerrero A., Milosevic I., and Raimundo N.. . 2019. Mitochondrial fission requires DRP1 but not dynamins . Nature 570 :E34–E42. doi: 10.1038/s41586-019-1296-y. [PubMed] [CrossRef] [Google Scholar]
- Fu, L., Xu Y., Hou Y., Qi X., Zhou L., Liu H., Luan Y., Jing L., Miao Y., Zhao S., . et al. 2017. Proteomic analysis indicates that mitochondrial energy metabolism in skeletal muscle tissue is negatively correlated with feed efficiency in pigs . Sci. Rep . 7 :45291. doi: 10.1038/srep45291. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Fu, X., Zhu M. J., Dodson M. V., and Du M.. . 2015. AMP-activated protein kinase stimulates Warburg-like glycolysis and activation of satellite cells during muscle regeneration . J. Biol. Chem . 290 :26445–26456. doi: 10.1074/jbc.M115.665232. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Fu, X., Zhu M., Zhang S., Foretz M., Viollet B., and Du M.. . 2016. Obesity impairs skeletal muscle regeneration through inhibition of AMPK . Diabetes 65 :188–200. doi: 10.2337/db15-0647. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Fueger, P. T., Bracy D. P., Malabanan C. M., Pencek R. R., Granner D. K., and Wasserman D. H.. . 2004. Hexokinase II overexpression improves exercise-stimulated but not insulin-stimulated muscle glucose uptake in high-fat-fed C57BL/6J mice . Diabetes 53 :306–314. doi: 10.2337/diabetes.53.2.306. [PubMed] [CrossRef] [Google Scholar]
- Garcia-Roves, P. M., Osler M. E., Holmström M. H., and Zierath J. R.. . 2008. Gain-of-function R225Q mutation in AMP-activated protein kinase gamma3 subunit increases mitochondrial biogenesis in glycolytic skeletal muscle . J. Biol. Chem . 283 :35724–35734. doi: 10.1074/jbc.M805078200. [PubMed] [CrossRef] [Google Scholar]
- Ghelli, A., Tropeano C. V., Calvaruso M. A., Marchesini A., Iommarini L., Porcelli A. M., Zanna C., De Nardo V., Martinuzzi A., Wibrand F., . et al. 2013. The cytochrome b p.278Y>C mutation causative of a multisystem disorder enhances superoxide production and alters supramolecular interactions of respiratory chain complexes . Hum. Mol. Genet . 22 :2141–2151. doi: 10.1093/hmg/ddt067. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Glancy, B., Hartnell L. M., Malide D., Yu Z. X., Combs C. A., Connelly P. S., Subramaniam S., and Balaban R. S.. . 2015. Mitochondrial reticulum for cellular energy distribution in muscle . Nature 523 :617–620. doi: 10.1038/nature14614. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Glancy, B., Hsu L. Y., Dao L., Bakalar M., French S., Chess D. J., Taylor J. L., Picard M., Aponte A., Daniels M. P., . et al. 2014. In vivo microscopy reveals extensive embedding of capillaries within the sarcolemma of skeletal muscle fibers . Microcirculation 21 :131–147. doi: 10.1111/micc.12098. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Goldstein, J. C., Waterhouse N. J., Juin P., Evan G. I., and Green D. R.. . 2000. The coordinate release of cytochrome c during apoptosis is rapid, complete and kinetically invariant . Nat. Cell Biol . 2 :156–162. doi: 10.1038/35004029. [PubMed] [CrossRef] [Google Scholar]
- Gomes, L. C., Di Benedetto G., and Scorrano L.. . 2011. During autophagy mitochondria elongate, are spared from degradation and sustain cell viability . Nat. Cell Biol . 13 :589–598. doi: 10.1038/ncb2220. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Greggio, C., Jha P., Kulkarni S. S., Lagarrigue S., Broskey N. T., Boutant M., Wang X., Conde Alonso S., Ofori E., Auwerx J., . et al. 2017. Enhanced respiratory chain supercomplex formation in response to exercise in human skeletal muscle . Cell Metab . 25 :301–311. doi: 10.1016/j.cmet.2016.11.004. [PubMed] [CrossRef] [Google Scholar]
- Grobet, L., Martin L. J., Poncelet D., Pirottin D., Brouwers B., Riquet J., Schoeberlein A., Dunner S., Ménissier F., Massabanda J., . et al. 1997. A deletion in the bovine myostatin gene causes the double-muscled phenotype in cattle . Nat. Genet . 17 :71–74. doi: 10.1038/ng0997-71. [PubMed] [CrossRef] [Google Scholar]
- Grzelkowska-Kowalczyk, K., Wieteska-Skrzeczyńska W., Grabiec K., and Tokarska J.. . 2013. High glucose-mediated alterations of mechanisms important in myogenesis of mouse C2C12 myoblasts . Cell Biol. Int . 37 :29–35. doi: 10.1002/cbin.10004. [PubMed] [CrossRef] [Google Scholar]
- Gunawan, A. M., Richert B. T., Schinckel A. P., Grant A. L., and Gerrard D. E.. . 2007. Ractopamine induces differential gene expression in porcine skeletal muscles . J. Anim. Sci . 85 :2115–2124. doi: 10.2527/jas.2006-540. [PubMed] [CrossRef] [Google Scholar]
- Gunawan, A. M., Yen C. N., Richert B. T., Schinckel A. P., Grant A. L., and Gerrard D. E.. . 2020. Ractopamine-induced fiber type-specific gene expression in porcine skeletal muscles is independent of growth . J. Anim. Sci . 98 ( 11 ):1–16. doi: 10.1093/jas/skaa341. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Hamai, N., Nakamura M., and Asano A.. . 1997. Inhibition of mitochondrial protein synthesis impaired C2C12 myoblast differentiation . Cell Struct. Funct . 22 :421–431. doi: 10.1247/csf.22.421. [PubMed] [CrossRef] [Google Scholar]
- Hamrick, M. W., Pennington C., Webb C. N., and Isales C. M.. . 2006. Resistance to body fat gain in ‘double-muscled’ mice fed a high-fat diet . Int. J. Obes. (Lond) 30 :868–870. doi: 10.1038/sj.ijo.0803200. [PubMed] [CrossRef] [Google Scholar]
- Han, D. H., Hansen P. A., Host H. H., and Holloszy J. O.. . 1997. Insulin resistance of muscle glucose transport in rats fed a high-fat diet: a reevaluation . Diabetes 46 :1761–1767. doi: 10.2337/diab.46.11.1761. [PubMed] [CrossRef] [Google Scholar]
- Han, B., Tong J., Zhu M. J., Ma C., and Du M.. . 2008. Insulin-like growth factor-1 (IGF-1) and leucine activate pig myogenic satellite cells through mammalian target of rapamycin (mTOR) pathway . Mol. Reprod. Dev . 75 :810–817. doi: 10.1002/mrd.20832. [PubMed] [CrossRef] [Google Scholar]
- Handschin, C., Chin S., Li P., Liu F., Maratos-Flier E., Lebrasseur N. K., Yan Z., and Spiegelman B. M.. . 2007. Skeletal muscle fiber-type switching, exercise intolerance, and myopathy in PGC-1 alpha muscle-specific knock-out animals . J. Biol. Chem . 282 :30014–30021. doi: 10.1074/jbc.M704817200. [PubMed] [CrossRef] [Google Scholar]
- Hanover, J. A., Krause M. W., and Love D. C.. . 2012. Bittersweet memories: linking metabolism to epigenetics through O-GlcNAcylation . Nat. Rev. Mol. Cell Biol . 13 :312–321. doi: 10.1038/nrm3334. [PubMed] [CrossRef] [Google Scholar]
- Hanover, J. A., Lai Z., Lee G., Lubas W. A., and Sato S. M.. . 1999. Elevated O-linked N-acetylglucosamine metabolism in pancreatic beta-cells . Arch. Biochem. Biophys . 362 :38–45. doi: 10.1006/abbi.1998.1016. [PubMed] [CrossRef] [Google Scholar]
- Hansford, R. G. 1994. Physiological role of mitochondrial Ca2 + transport . J. Bioenerg. Biomembr . 26 :495–508. doi: 10.1007/BF00762734. [PubMed] [CrossRef] [Google Scholar]
- Hansford, R. G., and Zorov D.. . 1998. Role of mitochondrial calcium transport in the control of substrate oxidation . Mol. Cell. Biochem . 184 :359–369. doi: 10.1023/A:1006893903113. [PubMed] [CrossRef] [Google Scholar]
- Harwood, K. R., and Hanover J. A.. . 2014. Nutrient-driven O-GlcNAc cycling—think globally but act locally . J. Cell Sci . 127 ( Pt 9 ):1857–1867. doi: 10.1242/jcs.113233. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Hasty, P., Bradley A., Morris J. H., Edmondson D. G., Venuti J. M., Olson E. N., and Klein W. H.. . 1993. Muscle deficiency and neonatal death in mice with a targeted mutation in the myogenin gene . Nature 364 :501–506. doi: 10.1038/364501a0. [PubMed] [CrossRef] [Google Scholar]
- Hatfield, I., Harvey I., Yates E. R., Redd J. R., Reiter L. T., and Bridges D.. . 2015. The role of TORC1 in muscle development in Drosophila . Sci. Rep . 5 :9676. doi: 10.1038/srep09676. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Hawkins, M., Angelov I., Liu R., Barzilai N., and Rossetti L.. . 1997. The tissue concentration of UDP-N-acetylglucosamine modulates the stimulatory effect of insulin on skeletal muscle glucose uptake . J. Biol. Chem . 272 :4889–4895. doi: 10.1074/jbc.272.8.4889. [PubMed] [CrossRef] [Google Scholar]
- Hawkins, M., Barzilai N., Liu R., Hu M., Chen W., and Rossetti L.. . 1997. Role of the glucosamine pathway in fat-induced insulin resistance . J. Clin. Invest . 99 :2173–2182. doi: 10.1172/JCI119390. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Hayashi, T., Wojtaszewski J. F., and Goodyear L. J.. . 1997. Exercise regulation of glucose transport in skeletal muscle . Am. J. Physiol . 273 :E1039–E1051. doi: 10.1152/ajpendo.1997.273.6.E1039. [PubMed] [CrossRef] [Google Scholar]
- Heineman, F. W., and Balaban R. S.. . 1990. Phosphorus-31 nuclear magnetic resonance analysis of transient changes of canine myocardial metabolism in vivo . J. Clin. Invest . 85 :843–852. doi: 10.1172/JCI114511. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Herzberg, N. H., Zwart R., Wolterman R. A., Ruiter J. P., Wanders R. J., Bolhuis P. A., and van den Bogert C.. . 1993. Differentiation and proliferation of respiration-deficient human myoblasts . Biochim. Biophys. Acta 1181 :63–67. doi: 10.1016/0925-4439(93)90091-e. [PubMed] [CrossRef] [Google Scholar]
- Herzig, S., Raemy E., Montessuit S., Veuthey J. L., Zamboni N., Westermann B., Kunji E. R., and Martinou J. C.. . 2012. Identification and functional expression of the mitochondrial pyruvate carrier . Science 337 :93–96. doi: 10.1126/science.1218530. [PubMed] [CrossRef] [Google Scholar]
- Hinson, R. B., Wiegand B. R., Ritter M. J., Allee G. L., and Carr S. N.. . 2011. Impact of dietary energy level and ractopamine on growth performance, carcass characteristics, and meat quality of finishing pigs . J. Anim. Sci . 89 :3572–3579. doi: 10.2527/jas.2010-3302. [PubMed] [CrossRef] [Google Scholar]
- Holloszy, J. O. 1967. Biochemical adaptations in muscle. Effects of exercise on mitochondrial oxygen uptake and respiratory enzyme activity in skeletal muscle . J. Biol. Chem . 242 :2278–2282. doi: 10.1016/S0021-9258(18)96046-1. [PubMed] [CrossRef] [Google Scholar]
- Holness, M. J., Kraus A., Harris R. A., and Sugden M. C.. . 2000. Targeted upregulation of pyruvate dehydrogenase kinase (PDK)-4 in slow-twitch skeletal muscle underlies the stable modification of the regulatory characteristics of PDK induced by high-fat feeding . Diabetes 49 :775–781. doi: 10.2337/diabetes.49.5.775. [PubMed] [CrossRef] [Google Scholar]
- Hori, S., Hiramuki Y., Nishimura D., Sato F., and Sehara-Fujisawa A.. . 2019. PDH-mediated metabolic flow is critical for skeletal muscle stem cell differentiation and myotube formation during regeneration in mice . FASEB J . 33 :8094–8109. doi: 10.1096/fj.201802479R. [PubMed] [CrossRef] [Google Scholar]
- Housley, M. P., Rodgers J. T., Udeshi N. D., Kelly T. J., Shabanowitz J., Hunt D. F., Puigserver P., and Hart G. W.. . 2008. O-GlcNAc regulates FoxO activation in response to glucose . J. Biol. Chem . 283 :16283–16292. doi: 10.1074/jbc.M802240200. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Hu, Y., Suarez J., Fricovsky E., Wang H., Scott B. T., Trauger S. A., Han W., Hu Y., Oyeleye M. O., and Dillmann W. H.. . 2009. Increased enzymatic O-GlcNAcylation of mitochondrial proteins impairs mitochondrial function in cardiac myocytes exposed to high glucose . J. Biol. Chem . 284 :547–555. doi: 10.1074/jbc.M808518200. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Huizing, M., Ruitenbeek W., Thinnes F. P., DePinto V., Wendel U., Trijbels F. J., Smit L. M., ter Laak H. J., and van den Heuvel L. P.. . 1996. Deficiency of the voltage-dependent anion channel: a novel cause of mitochondriopathy . Pediatr. Res . 39 :760–765. doi: 10.1203/00006450-199605000-00003. [PubMed] [CrossRef] [Google Scholar]
- Izumiya, Y., Hopkins T., Morris C., Sato K., Zeng L., Viereck J., Hamilton J. A., Ouchi N., LeBrasseur N. K., and Walsh K.. . 2008. Fast/glycolytic muscle fiber growth reduces fat mass and improves metabolic parameters in obese mice . Cell Metab . 7 :159–172. doi: 10.1016/j.cmet.2007.11.003. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Jaleel, A., Short K. R., Asmann Y. W., Klaus K. A., Morse D. M., Ford G. C., and Nair K. S.. . 2008. In vivo measurement of synthesis rate of individual skeletal muscle mitochondrial proteins . Am. J. Physiol. Endocrinol. Metab . 295 :E1255–E1268. doi: 10.1152/ajpendo.90586.2008. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Jeoung, N. H., and Harris R. A.. . 2008. Pyruvate dehydrogenase kinase-4 deficiency lowers blood glucose and improves glucose tolerance in diet-induced obese mice . Am. J. Physiol. Endocrinol. Metab . 295 :E46–E54. doi: 10.1152/ajpendo.00536.2007. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Jeoung, N. H., Wu P., Joshi M. A., Jaskiewicz J., Bock C. B., Depaoli-Roach A. A., and Harris R. A.. . 2006. Role of pyruvate dehydrogenase kinase isoenzyme 4 (PDHK4) in glucose homoeostasis during starvation . Biochem. J . 397 :417–425. doi: 10.1042/BJ20060125. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Jessen, N., and Goodyear L. J.. . 2005. Contraction signaling to glucose transport in skeletal muscle . j. Appl. Physiol. (1985) . 99 :330–337. doi: 10.1152/japplphysiol.00175.2005. [PubMed] [CrossRef] [Google Scholar]
- Ji, W. K., Hatch A. L., Merrill R. A., Strack S., and Higgs H. N.. . 2015. Actin filaments target the oligomeric maturation of the dynamin GTPase Drp1 to mitochondrial fission sites . Elife 4 :e11553. doi: 10.7554/eLife.11553. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Jin, C. L., Ye J. L., Yang J., Gao C. Q., Yan H. C., Li H. C., and Wang X. Q.. . 2019. mTORC1 mediates lysine-induced satellite cell activation to promote skeletal muscle growth . Cells 8 ( 12 ). doi: 10.3390/cells8121549. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Jing, L., Hou Y., Wu H., Miao Y., Li X., Cao J., Brameld J. M., Parr T., and Zhao S.. . 2015. Transcriptome analysis of mRNA and miRNA in skeletal muscle indicates an important network for differential residual feed intake in pigs . Sci. Rep . 5 :11953. doi: 10.1038/srep11953. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Kalia, R., Wang R. Y., Yusuf A., Thomas P. V., Agard D. A., Shaw J. M., and Frost A.. . 2018. Structural basis of mitochondrial receptor binding and constriction by DRP1 . Nature 558 :401–405. doi: 10.1038/s41586-018-0211-2. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Kalista, S., Schakman O., Gilson H., Lause P., Demeulder B., Bertrand L., Pende M., and Thissen J. P.. . 2012. The type 1 insulin-like growth factor receptor (IGF-IR) pathway is mandatory for the follistatin-induced skeletal muscle hypertrophy . Endocrinology 153 :241–253. doi: 10.1210/en.2011-1687. [PubMed] [CrossRef] [Google Scholar]
- Kang, E. S., Han D., Park J., Kwak T. K., Oh M. A., Lee S. A., Choi S., Park Z. Y., Kim Y., and Lee J. W.. . 2008. O-GlcNAc modulation at Akt1 Ser473 correlates with apoptosis of murine pancreatic beta cells . Exp. Cell Res . 314 :2238–2248. doi: 10.1016/j.yexcr.2008.04.014. [PubMed] [CrossRef] [Google Scholar]
- Katz, A., Broberg S., Sahlin K., and Wahren J.. . 1986. Leg glucose uptake during maximal dynamic exercise in humans . Am. J. Physiol . 251 ( 1 Pt 1 ):E65–E70. doi: 10.1152/ajpendo.1986.251.1.E65. [PubMed] [CrossRef] [Google Scholar]
- Kennedy, J. W., Hirshman M. F., Gervino E. V., Ocel J. V., Forse R. A., Hoenig S. J., Aronson D., Goodyear L. J., and Horton E. S.. . 1999. Acute exercise induces GLUT4 translocation in skeletal muscle of normal human subjects and subjects with type 2 diabetes . Diabetes 48 :1192–1197. doi: 10.2337/diabetes.48.5.1192. [PubMed] [CrossRef] [Google Scholar]
- Kim, J., Kundu M., Viollet B., and Guan K. L.. . 2011. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1 . Nat. Cell Biol . 13 :132–141. doi: 10.1038/ncb2152. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Korohoda, W., Pietrzkowski Z., and Reiss K.. . 1993. Chloramphenicol, an inhibitor of mitochondrial protein synthesis, inhibits myoblast fusion and myotube differentiation . Folia Histochem. Cytobiol . 31 :9–13. [PubMed] [Google Scholar]
- Korzeniewski, B. 2007. Regulation of oxidative phosphorylation through parallel activation . Biophys. Chem . 129 :93–110. doi: 10.1016/j.bpc.2007.05.013. [PubMed] [CrossRef] [Google Scholar]
- Koshiba, T., Detmer S. A., Kaiser J. T., Chen H., McCaffery J. M., and Chan D. C.. . 2004. Structural basis of mitochondrial tethering by mitofusin complexes . Science 305 :858–862. doi: 10.1126/science.1099793. [PubMed] [CrossRef] [Google Scholar]
- Kuang, S., Kuroda K., Le Grand F., and Rudnicki M. A.. . 2007. Asymmetric self-renewal and commitment of satellite stem cells in muscle . Cell 129 :999–1010. doi: 10.1016/j.cell.2007.03.044. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Kurth-Kraczek, E. J., Hirshman M. F., Goodyear L. J., and Winder W. W.. . 1999. 5’ AMP-activated protein kinase activation causes GLUT4 translocation in skeletal muscle . Diabetes 48 :1667–1671. doi: 10.2337/diabetes.48.8.1667. [PubMed] [CrossRef] [Google Scholar]
- Lai, K. M., Gonzalez M., Poueymirou W. T., Kline W. O., Na E., Zlotchenko E., Stitt T. N., Economides A. N., Yancopoulos G. D., and Glass D. J.. . 2004. Conditional activation of AKT in adult skeletal muscle induces rapid hypertrophy . Mol. Cell. Biol . 24 :9295–9304. doi: 10.1128/MCB.24.21.9295-9304.2004. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Lantier, L., Mounier R., Leclerc J., Pende M., Foretz M., and Viollet B.. . 2010. Coordinated maintenance of muscle cell size control by AMP-activated protein kinase . FASEB J . 24 :3555–3561. doi: 10.1096/fj.10-155994. [PubMed] [CrossRef] [Google Scholar]
- Le Naou, T., Le Floc’h N., Louveau I., Gilbert H., and Gondret F.. . 2012. Metabolic changes and tissue responses to selection on residual feed intake in growing pigs . J. Anim. Sci . 90 :4771–4780. doi: 10.2527/jas.2012-5226. [PubMed] [CrossRef] [Google Scholar]
- Lee, J. W., Park S., Takahashi Y., and Wang H. G.. . 2010. The association of AMPK with ULK1 regulates autophagy . PLoS One 5 :e15394. doi: 10.1371/journal.pone.0015394. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Lepper, C., Partridge T. A., and Fan C. M.. . 2011. An absolute requirement for Pax7-positive satellite cells in acute injury-induced skeletal muscle regeneration . Development 138 :3639–3646. doi: 10.1242/dev.067595. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Lewis, S. E., Kelly F. J., and Goldspink D. F.. . 1984. Pre- and post-natal growth and protein turnover in smooth muscle, heart and slow- and fast-twitch skeletal muscles of the rat . Biochem. J . 217 :517–526. doi: 10.1042/bj2170517. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Li, P., Zeng Z., Wang D., Xue L., Zhang R., and Piao X.. . 2012. Effects of the standardized ileal digestible lysine to metabolizable energy ratio on performance and carcass characteristics of growing-finishing pigs . J. Anim. Sci. Biotechnol . 3 :9. doi: 10.1186/2049-1891-3-9. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Lipina, C., Kendall H., McPherron A. C., Taylor P. M., and Hundal H. S.. . 2010. Mechanisms involved in the enhancement of mammalian target of rapamycin signalling and hypertrophy in skeletal muscle of myostatin-deficient mice . FEBS Lett . 584 :2403–2408. doi: 10.1016/j.febslet.2010.04.039. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Ljubicic, V., Miura P., Burt M., Boudreault L., Khogali S., Lunde J. A., Renaud J. M., and Jasmin B. J.. . 2011. Chronic AMPK activation evokes the slow, oxidative myogenic program and triggers beneficial adaptations in mdx mouse skeletal muscle . Hum. Mol. Genet . 20 :3478–3493. doi: 10.1093/hmg/ddr265. [PubMed] [CrossRef] [Google Scholar]
- Lunt, S. Y., and Vander Heiden M. G.. . 2011. Aerobic glycolysis: meeting the metabolic requirements of cell proliferation . Annu. Rev. Cell Dev. Biol . 27 :441–464. doi: 10.1146/annurev-cellbio-092910-154237. [PubMed] [CrossRef] [Google Scholar]
- Luo, W., Ai L., Wang B. F., and Zhou Y.. . 2019. High glucose inhibits myogenesis and induces insulin resistance by down-regulating AKT signaling . Biomed. Pharmacother . 120 :109498. doi: 10.1016/j.biopha.2019.109498. [PubMed] [CrossRef] [Google Scholar]
- Main, R., Dritz S. S., Tokach M. D., Goodband R. D., Nelssen J. L., and DeRouchey J. M.. . 2009. Effects of ractopamine HCl dose and treatment period on pig performance in a commercial finishing facility . Journal of Swine Health and Production 17 ( 3 ):134–139. [Google Scholar]
- Maldonado, E. N., and Lemasters J. J.. . 2014. ATP/ADP ratio, the missed connection between mitochondria and the Warburg effect . Mitochondrion 19 Pt A :78–84. doi: 10.1016/j.mito.2014.09.002. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Maranzana, E., Barbero G., Falasca A. I., Lenaz G., and Genova M. L.. . 2013. Mitochondrial respiratory supercomplex association limits production of reactive oxygen species from complex I . Antioxid Redox Signal . 19 :1469–1480. doi: 10.1089/ars.2012.4845. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Marshall, S., Bacote V., and Traxinger R. R.. . 1991. Discovery of a metabolic pathway mediating glucose-induced desensitization of the glucose transport system. Role of hexosamine biosynthesis in the induction of insulin resistance . J. Biol. Chem . 266 :4706–4712. [PubMed] [Google Scholar]
- Martínez-Reyes, I., and Chandel N. S.. . 2020. Mitochondrial TCA cycle metabolites control physiology and disease . Nat. Commun . 11 :102. doi: 10.1038/s41467-019-13668-3. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Mauro, A. 1961. Satellite cell of skeletal muscle fibers . J. Biophys. Biochem. Cytol . 9 :493–495. doi: 10.1083/jcb.9.2.493. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Mavalli, M. D., DiGirolamo D. J., Fan Y., Riddle R. C., Campbell K. S., van Groen T., Frank S. J., Sperling M. A., Esser K. A., Bamman M. M., . et al. 2010. Distinct growth hormone receptor signaling modes regulate skeletal muscle development and insulin sensitivity in mice . J. Clin. Invest . 120 :4007–4020. doi: 10.1172/JCI42447. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- McCormack, J. G., and Denton R. M.. . 1979. The effects of calcium ions and adenine nucleotides on the activity of pig heart 2-oxoglutarate dehydrogenase complex . Biochem. J . 180 :533–544. doi: 10.1042/bj1800533. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- McCormack, J. G., and Denton R. M.. . 1993. Mitochondrial Ca2+ transport and the role of intramitochondrial Ca2+ in the regulation of energy metabolism . Dev. Neurosci . 15 :165–173. doi: 10.1159/000111332. [PubMed] [CrossRef] [Google Scholar]
- McCormack, J. G., Halestrap A. P., and Denton R. M.. . 1990. Role of calcium ions in regulation of mammalian intramitochondrial metabolism . Physiol. Rev . 70 :391–425. doi: 10.1152/physrev.1990.70.2.391. [PubMed] [CrossRef] [Google Scholar]
- McPherron, A. C., and Lee S. J.. . 1997. Double muscling in cattle due to mutations in the myostatin gene . Proc. Natl. Acad. Sci. U.S.A . 94 :12457–12461. doi: 10.1073/pnas.94.23.12457. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- McPherron, A. C., and Lee S. J.. . 2002. Suppression of body fat accumulation in myostatin-deficient mice . J. Clin. Invest . 109 :595–601. doi: 10.1172/JCI13562. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Meeusen, S., DeVay R., Block J., Cassidy-Stone A., Wayson S., McCaffery J. M., and Nunnari J.. . 2006. Mitochondrial inner-membrane fusion and crista maintenance requires the dynamin-related GTPase Mgm1 . Cell 127 :383–395. doi: 10.1016/j.cell.2006.09.021. [PubMed] [CrossRef] [Google Scholar]
- Meeusen, S., McCaffery J. M., and Nunnari J.. . 2004. Mitochondrial fusion intermediates revealed in vitro . Science 305 :1747–1752. doi: 10.1126/science.1100612. [PubMed] [CrossRef] [Google Scholar]
- Meng, Z. X., Li S., Wang L., Ko H. J., Lee Y., Jung D. Y., Okutsu M., Yan Z., Kim J. K., and Lin J. D.. . 2013. Baf60c drives glycolytic metabolism in the muscle and improves systemic glucose homeostasis through Deptor-mediated AKT activation . Nat. Med . 19 :640–645. doi: 10.1038/nm.3144. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Menshikova, E. V., Ritov V. B., Fairfull L., Ferrell R. E., Kelley D. E., and Goodpaster B. H.. . 2006. Effects of exercise on mitochondrial content and function in aging human skeletal muscle . J. Gerontol. A Biol. Sci. Med. Sci . 61 :534–540. doi: 10.1093/gerona/61.6.534. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Mishra, P., Varuzhanyan G., Pham A. H., and Chan D. C.. . 2015. Mitochondrial dynamics is a distinguishing feature of skeletal muscle fiber types and regulates organellar compartmentalization . Cell Metab . 22 :1033–1044. doi: 10.1016/j.cmet.2015.09.027. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Mootha, V. K., Lindgren C. M., Eriksson K. F., Subramanian A., Sihag S., Lehar J., Puigserver P., Carlsson E., Ridderstråle M., Laurila E., . et al. 2003. PGC-1alpha-responsive genes involved in oxidative phosphorylation are coordinately downregulated in human diabetes . Nat. Genet . 34 :267–273. doi: 10.1038/ng1180. [PubMed] [CrossRef] [Google Scholar]
- Mosher, D. S., Quignon P., Bustamante C. D., Sutter N. B., Mellersh C. S., Parker H. G., and Ostrander E. A.. . 2007. A mutation in the myostatin gene increases muscle mass and enhances racing performance in heterozygote dogs . PLoS Genet . 3 :e79. doi: 10.1371/journal.pgen.0030079. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Moss, F. P., and Leblond C. P.. . 1971. Satellite cells as the source of nuclei in muscles of growing rats . Anat. Rec . 170 :421–435. doi: 10.1002/ar.1091700405. [PubMed] [CrossRef] [Google Scholar]
- Mouisel, E., Relizani K., Mille-Hamard L., Denis R., Hourdé C., Agbulut O., Patel K., Arandel L., Morales-Gonzalez S., Vignaud A., . et al. 2014. Myostatin is a key mediator between energy metabolism and endurance capacity of skeletal muscle . Am. J. Physiol. Regul. Integr. Comp. Physiol . 307 :R444–R454. doi: 10.1152/ajpregu.00377.2013. [PubMed] [CrossRef] [Google Scholar]
- Moyes, C. D., Mathieu-Costello O. A., Tsuchiya N., Filburn C., and Hansford R. G.. . 1997. Mitochondrial biogenesis during cellular differentiation . Am. J. Physiol . 272 ( 4 Pt 1 ):C1345–C1351. doi: 10.1152/ajpcell.1997.272.4.C1345. [PubMed] [CrossRef] [Google Scholar]
- Musarò, A., McCullagh K., Paul A., Houghton L., Dobrowolny G., Molinaro M., Barton E. R., Sweeney H. L., and Rosenthal N.. . 2001. Localized Igf-1 transgene expression sustains hypertrophy and regeneration in senescent skeletal muscle . Nat. Genet . 27 :195–200. doi: 10.1038/84839. [PubMed] [CrossRef] [Google Scholar]
- Nabeshima, Y., Hanaoka K., Hayasaka M., Esumi E., Li S., Nonaka I., and Nabeshima Y.. . 1993. Myogenin gene disruption results in perinatal lethality because of severe muscle defect . Nature 364 :532–535. doi: 10.1038/364532a0. [PubMed] [CrossRef] [Google Scholar]
- Nuth, M., Yoon T., and Cowan J. A.. . 2002. Iron-sulfur cluster biosynthesis: characterization of iron nucleation sites for assembly of the [2Fe-2S]2+ cluster core in IscU proteins . J. Am. Chem. Soc . 124 :8774–8775. doi: 10.1021/ja0264596. [PubMed] [CrossRef] [Google Scholar]
- Palmieri, F., and Pierri C. L.. . 2010. Mitochondrial metabolite transport . Essays Biochem . 47 :37–52. doi: 10.1042/bse0470037. [PubMed] [CrossRef] [Google Scholar]
- Patti, M. E., Butte A. J., Crunkhorn S., Cusi K., Berria R., Kashyap S., Miyazaki Y., Kohane I., Costello M., Saccone R., . et al. 2003. Coordinated reduction of genes of oxidative metabolism in humans with insulin resistance and diabetes: potential role of PGC1 and NRF1 . Proc. Natl. Acad. Sci. U.s.a . 100 :8466–8471. doi: 10.1073/pnas.1032913100. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Pereira, M. G., Baptista I. L., Carlassara E. O., Moriscot A. S., Aoki M. S., and Miyabara E. H.. . 2014. Leucine supplementation improves skeletal muscle regeneration after cryolesion in rats . PLoS One 9 :e85283. doi: 10.1371/journal.pone.0085283. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Peterson, J. M., Bryner R. W., and Alway S. E.. . 2008. Satellite cell proliferation is reduced in muscles of obese Zucker rats but restored with loading . Am. J. Physiol. Cell Physiol . 295 :C521–C528. doi: 10.1152/ajpcell.00073.2008. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Pfeiffer, T., Schuster S., and Bonhoeffer S.. . 2001. Cooperation and competition in the evolution of ATP-producing pathways . Science 292 :504–507. doi: 10.1126/science.1058079. [PubMed] [CrossRef] [Google Scholar]
- Picard, M., White K., and Turnbull D. M.. . 2013. Mitochondrial morphology, topology, and membrane interactions in skeletal muscle: a quantitative three-dimensional electron microscopy study . J. Appl. Physiol. (1985) . 114 :161–171. doi: 10.1152/japplphysiol.01096.2012. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Rao, X., Duan X., Mao W., Li X., Li Z., Li Q., Zheng Z., Xu H., Chen M., Wang P. G., . et al. 2015. O-GlcNAcylation of G6PD promotes the pentose phosphate pathway and tumor growth . Nat. Commun . 6 :8468. doi: 10.1038/ncomms9468. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Reisz-Porszasz, S., Bhasin S., Artaza J. N., Shen R., Sinha-Hikim I., Hogue A., Fielder T. J., and Gonzalez-Cadavid N. F.. . 2003. Lower skeletal muscle mass in male transgenic mice with muscle-specific overexpression of myostatin . Am. J. Physiol. Endocrinol. Metab . 285 :E876–E888. doi: 10.1152/ajpendo.00107.2003. [PubMed] [CrossRef] [Google Scholar]
- Remels, A. H., Langen R. C., Schrauwen P., Schaart G., Schols A. M., and Gosker H. R.. . 2010. Regulation of mitochondrial biogenesis during myogenesis . Mol. Cell. Endocrinol . 315 :113–120. doi: 10.1016/j.mce.2009.09.029. [PubMed] [CrossRef] [Google Scholar]
- Risson, V., Mazelin L., Roceri M., Sanchez H., Moncollin V., Corneloup C., Richard-Bulteau H., Vignaud A., Baas D., Defour A., . et al. 2009. Muscle inactivation of mTOR causes metabolic and dystrophin defects leading to severe myopathy . J. Cell Biol . 187 :859–874. doi: 10.1083/jcb.200903131. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Rochard, P., Rodier A., Casas F., Cassar-Malek I., Marchal-Victorion S., Daury L., Wrutniak C., and Cabello G.. . 2000. Mitochondrial activity is involved in the regulation of myoblast differentiation through myogenin expression and activity of myogenic factors . J. Biol. Chem . 275 :2733–2744. doi: 10.1074/jbc.275.4.2733. [PubMed] [CrossRef] [Google Scholar]
- Rocheteau, P., Gayraud-Morel B., Siegl-Cachedenier I., Blasco M. A., and Tajbakhsh S.. . 2012. A subpopulation of adult skeletal muscle stem cells retains all template DNA strands after cell division . Cell 148 :112–125. doi: 10.1016/j.cell.2011.11.049. [PubMed] [CrossRef] [Google Scholar]
- Rosholt, M. N., King P. A., and Horton E. S.. . 1994. High-fat diet reduces glucose transporter responses to both insulin and exercise . Am. J. Physiol . 266 ( 1 Pt 2 ):R95–101. doi: 10.1152/ajpregu.1994.266.1.R95. [PubMed] [CrossRef] [Google Scholar]
- Rottenberg, H., and Scarpa A.. . 1974. Calcium uptake and membrane potential in mitochondria . Biochemistry 13 :4811–4817. doi: 10.1021/bi00720a020. [PubMed] [CrossRef] [Google Scholar]
- Ruan, H. B., Singh J. P., Li M. D., Wu J., and Yang X.. . 2013. Cracking the O-GlcNAc code in metabolism . Trends Endocrinol. Metab . 24 :301–309. doi: 10.1016/j.tem.2013.02.002. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Rudnicki, M. A., Braun T., Hinuma S., and Jaenisch R.. . 1992. Inactivation of MyoD in mice leads to up-regulation of the myogenic HLH gene Myf-5 and results in apparently normal muscle development . Cell 71 :383–390. doi: 10.1016/0092-8674(92)90508-a. [PubMed] [CrossRef] [Google Scholar]
- Sacoman, J. L., Dagda R. Y., Burnham-Marusich A. R., Dagda R. K., and Berninsone P. M.. . 2017. Mitochondrial O-GlcNAc transferase (mOGT) regulates mitochondrial structure, function, and survival in HeLa cells . J. Biol. Chem . 292 :4499–4518. doi: 10.1074/jbc.M116.726752. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Sahlin, K., Tonkonogi M., and Söderlund K.. . 1998. Energy supply and muscle fatigue in humans . Acta Physiol. Scand . 162 :261–266. doi: 10.1046/j.1365-201X.1998.0298f.x. [PubMed] [CrossRef] [Google Scholar]
- Sato, T., Ito Y., Nedachi T., and Nagasawa T.. . 2014. Lysine suppresses protein degradation through autophagic-lysosomal system in C2C12 myotubes . Mol. Cell. Biochem . 391 :37–46. doi: 10.1007/s11010-014-1984-8. [PubMed] [CrossRef] [Google Scholar]
- Schägger, H., and Pfeiffer K.. . 2001. The ratio of oxidative phosphorylation complexes I-V in bovine heart mitochondria and the composition of respiratory chain supercomplexes . J. Biol. Chem . 276 :37861–37867. doi: 10.1074/jbc.M106474200. [PubMed] [CrossRef] [Google Scholar]
- Scheffler, T. L., Scheffler J. M., Park S., Kasten S. C., Wu Y., McMillan R. P., Hulver M. W., Frisard M. I., and Gerrard D. E.. . 2014. Fiber hypertrophy and increased oxidative capacity can occur simultaneously in pig glycolytic skeletal muscle . Am. J. Physiol. Cell Physiol . 306 :C354–C363. doi: 10.1152/ajpcell.00002.2013. [PubMed] [CrossRef] [Google Scholar]
- Schuelke, M., Wagner K. R., Stolz L. E., Hübner C., Riebel T., Kömen W., Braun T., Tobin J. F., and Lee S. J.. . 2004. Myostatin mutation associated with gross muscle hypertrophy in a child . N. Engl. J. Med . 350 :2682–2688. doi: 10.1056/NEJMoa040933. [PubMed] [CrossRef] [Google Scholar]
- Schuler, M., Ali F., Chambon C., Duteil D., Bornert J. M., Tardivel A., Desvergne B., Wahli W., Chambon P., and Metzger D.. . 2006. PGC1alpha expression is controlled in skeletal muscles by PPARbeta, whose ablation results in fiber-type switching, obesity, and type 2 diabetes . Cell Metab . 4 :407–414. doi: 10.1016/j.cmet.2006.10.003. [PubMed] [CrossRef] [Google Scholar]
- Schultz, E., Gibson M. C., and Champion T.. . 1978. Satellite cells are mitotically quiescent in mature mouse muscle: an EM and radioautographic study . J. Exp. Zool . 206 :451–456. doi: 10.1002/jez.1402060314. [PubMed] [CrossRef] [Google Scholar]
- Shi, H., Munk A., Nielsen T. S., Daughtry M. R., Larsson L., Li S., Høyer K. F., Geisler H. W., Sulek K., Kjøbsted R., . et al. 2018. Skeletal muscle O-GlcNAc transferase is important for muscle energy homeostasis and whole-body insulin sensitivity . Mol. Metab . 11 :160–177. doi: 10.1016/j.molmet.2018.02.010. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Short, K. R., Bigelow M. L., Kahl J., Singh R., Coenen-Schimke J., Raghavakaimal S., and Nair K. S.. . 2005. Decline in skeletal muscle mitochondrial function with aging in humans . Proc. Natl. Acad. Sci. U. S. A . 102 :5618–5623. doi: 10.1073/pnas.0501559102. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Shulman, G. I., Rothman D. L., Jue T., Stein P., DeFronzo R. A., and Shulman R. G.. . 1990. Quantitation of muscle glycogen synthesis in normal subjects and subjects with non-insulin-dependent diabetes by 13C nuclear magnetic resonance spectroscopy . N. Engl. J. Med . 322 :223–228. doi: 10.1056/NEJM199001253220403. [PubMed] [CrossRef] [Google Scholar]
- Simoneau, J. A., Colberg S. R., Thaete F. L., and Kelley D. E.. . 1995. Skeletal muscle glycolytic and oxidative enzyme capacities are determinants of insulin sensitivity and muscle composition in obese women . FASEB J . 9 :273–278. [PubMed] [Google Scholar]
- Skulachev, V. P. 1996. Role of uncoupled and non-coupled oxidations in maintenance of safely low levels of oxygen and its one-electron reductants . Q. Rev. Biophys . 29 :169–202. doi: 10.1017/s0033583500005795. [PubMed] [CrossRef] [Google Scholar]
- Slawson, C., Copeland R. J., and Hart G. W.. . 2010. O-GlcNAc signaling: a metabolic link between diabetes and cancer? Trends Biochem. Sci . 35 :547–555. doi: 10.1016/j.tibs.2010.04.005. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Snow, M. H. 1978. An autoradiographic study of satellite cell differentiation into regenerating myotubes following transplantation of muscles in young rats . Cell Tissue Res . 186 :535–540. doi: 10.1007/BF00224941. [PubMed] [CrossRef] [Google Scholar]
- Song, Z., Ghochani M., McCaffery J. M., Frey T. G., and Chan D. C.. . 2009. Mitofusins and OPA1 mediate sequential steps in mitochondrial membrane fusion . Mol. Biol. Cell 20 :3525–3532. doi: 10.1091/mbc.e09-03-0252. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Song, R., Peng W., Zhang Y., Lv F., Wu H. K., Guo J., Cao Y., Pi Y., Zhang X., Jin L., . et al. 2013. Central role of E3 ubiquitin ligase MG53 in insulin resistance and metabolic disorders . Nature 494 :375–379. doi: 10.1038/nature11834. [PubMed] [CrossRef] [Google Scholar]
- Stoppani, J., Hildebrandt A. L., Sakamoto K., Cameron-Smith D., Goodyear L. J., and Neufer P. D.. . 2002. AMP-activated protein kinase activates transcription of the UCP3 and HKII genes in rat skeletal muscle . Am. J. Physiol. Endocrinol. Metab . 283 :E1239–E1248. doi: 10.1152/ajpendo.00278.2002. [PubMed] [CrossRef] [Google Scholar]
- Swamy, M., Pathak S., Grzes K. M., Damerow S., Sinclair L. V., van Aalten D. M., and Cantrell D. A.. . 2016. Glucose and glutamine fuel protein O-GlcNAcylation to control T cell self-renewal and malignancy . Nat. Immunol . 17 :712–720. doi: 10.1038/ni.3439. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Tremblay, F., Lavigne C., Jacques H., and Marette A.. . 2001. Defective insulin-induced GLUT4 translocation in skeletal muscle of high fat-fed rats is associated with alterations in both Akt/protein kinase B and atypical protein kinase C (zeta/lambda) activities . Diabetes 50 :1901–1910. doi: 10.2337/diabetes.50.8.1901. [PubMed] [CrossRef] [Google Scholar]
- Tureckova, J., Wilson E. M., Cappalonga J. L., and Rotwein P.. . 2001. Insulin-like growth factor-mediated muscle differentiation: collaboration between phosphatidylinositol 3-kinase-Akt-signaling pathways and myogenin . J. Biol. Chem . 276 :39264–39270. doi: 10.1074/jbc.M104991200. [PubMed] [CrossRef] [Google Scholar]
- Uemichi, K., Shirai T., Hanakita H., and Takemasa T.. . 2021. Effect of mechanistic/mammalian target of rapamycin complex 1 on mitochondrial dynamics during skeletal muscle hypertrophy . Physiol. Rep . 9 :e14789. doi: 10.14814/phy2.14789. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Ustanina, S., Carvajal J., Rigby P., and Braun T.. . 2007. The myogenic factor Myf5 supports efficient skeletal muscle regeneration by enabling transient myoblast amplification . Stem Cells 25 :2006–2016. doi: 10.1634/stemcells.2006-0736. [PubMed] [CrossRef] [Google Scholar]
- Vasington, F. D., and Murphy J. V.. . 1962. Ca ion uptake by rat kidney mitochondria and its dependence on respiration and phosphorylation . J. Biol. Chem . 237 :2670–2677. [PubMed] [Google Scholar]
- Wagner, K. R., Kauffman F. C., and Max S. R.. . 1978. The pentose phosphate pathway in regenerating skeletal muscle . Biochem. J . 170 :17–22. doi: 10.1042/bj1700017. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Walgren, J. L., Vincent T. S., Schey K. L., and Buse M. G.. . 2003. High glucose and insulin promote O-GlcNAc modification of proteins, including alpha-tubulin . Am. J. Physiol. Endocrinol. Metab . 284 :E424–E434. doi: 10.1152/ajpendo.00382.2002. [PubMed] [CrossRef] [Google Scholar]
- Wang, Y. X., Zhang C. L., Yu R. T., Cho H. K., Nelson M. C., Bayuga-Ocampo C. R., Ham J., Kang H., and Evans R. M.. . 2004. Regulation of muscle fiber type and running endurance by PPARdelta . PLoS Biol . 2 :e294. doi: 10.1371/journal.pbio.0020294. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Wenz, T., Diaz F., Spiegelman B. M., and Moraes C. T.. . 2008. Activation of the PPAR/PGC-1alpha pathway prevents a bioenergetic deficit and effectively improves a mitochondrial myopathy phenotype . Cell Metab . 8 :249–256. doi: 10.1016/j.cmet.2008.07.006. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Retracted
- Wenz, T., Rossi S. G., Rotundo R. L., Spiegelman B. M., and Moraes C. T.. . 2009. Increased muscle PGC-1alpha expression protects from sarcopenia and metabolic disease during aging . Proc. Natl. Acad. Sci. U.S.A . 106 :20405–20410. doi: 10.1073/pnas.0911570106. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Retracted
- Williamson, D. L., Butler D. C., and Alway S. E.. . 2007. AMPK regulation of proliferation and differentiation in C2C12 culture models . FASEB J . 21 ( 6 ):A1205–A1206. doi: 10.1096/fasebj.21.6.A1205-d. [CrossRef] [Google Scholar]
- Williams, N. H., Cline T. R., Schinckel A. P., and Jones D. J.. . 1994. The impact of ractopamine, energy intake, and dietary fat on finisher pig growth performance and carcass merit . J. Anim. Sci . 72 :3152–3162. doi: 10.2527/1994.72123152x. [PubMed] [CrossRef] [Google Scholar]
- Willingham, T. B., Ajayi P. T., and Glancy B.. . 2021. Subcellular specialization of mitochondrial form and function in skeletal muscle cells . Front. Cell Dev. Biol . 9 :757305. doi: 10.3389/fcell.2021.757305. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Wu, C. L., Satomi Y., and Walsh K.. . 2017. RNA-seq and metabolomic analyses of Akt1-mediated muscle growth reveals regulation of regenerative pathways and changes in the muscle secretome . BMC Genomics 18 :181. doi: 10.1186/s12864-017-3548-2. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Wu, P., Sato J., Zhao Y., Jaskiewicz J., Popov K. M., and Harris R. A.. . 1998. Starvation and diabetes increase the amount of pyruvate dehydrogenase kinase isoenzyme 4 in rat heart . Biochem. J . 329 (Pt 1) :197–201. doi: 10.1042/bj3290197. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Yi, W., Clark P. M., Mason D. E., Keenan M. C., Hill C., W. A.Goddard, III, Peters E. C., Driggers E. M., and Hsieh-Wilson L. C.. . 2012. Phosphofructokinase 1 glycosylation regulates cell growth and metabolism . Science 337 :975–980. doi: 10.1126/science.1222278. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Yoon, T., and Cowan J. A.. . 2004. Frataxin-mediated iron delivery to ferrochelatase in the final step of heme biosynthesis . J. Biol. Chem . 279 :25943–25946. doi: 10.1074/jbc.C400107200. [PubMed] [CrossRef] [Google Scholar]
- Yousf, S., Sardesai D. M., Mathew A. B., Khandelwal R., Acharya J. D., Sharma S., and Chugh J.. . 2019. Metabolic signatures suggest o-phosphocholine to UDP-N-acetylglucosamine ratio as a potential biomarker for high-glucose and/or palmitate exposure in pancreatic β-cells . Metabolomics 15 :55. doi: 10.1007/s11306-019-1516-3. [PubMed] [CrossRef] [Google Scholar]
- Zeng, P. L., Yan H. C., Wang X. Q., Zhang C. M., Zhu C., Shu G., and Jiang Q. Y.. . 2013. Effects of dietary lysine levels on apparent nutrient digestibility and serum amino Acid absorption mode in growing pigs . Asian-Australas. J. Anim. Sci . 26 :1003–1011. doi: 10.5713/ajas.2012.12555. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Zhang, P., Liang X., Shan T., Jiang Q., Deng C., Zheng R., and Kuang S.. . 2015. mTOR is necessary for proper satellite cell activity and skeletal muscle regeneration . Biochem. Biophys. Res. Commun . 463 :102–108. doi: 10.1016/j.bbrc.2015.05.032. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
- Zheng, D., MacLean P. S., Pohnert S. C., Knight J. B., Olson A. L., Winder W. W., and Dohm G. L.. . 2001. Regulation of muscle GLUT-4 transcription by AMP-activated protein kinase . J. Appl. Physiol. (1985) . 91 :1073–1083. doi: 10.1152/jappl.2001.91.3.1073. [PubMed] [CrossRef] [Google Scholar]
- Zierath, J. R., Krook A., and Wallberg-Henriksson H.. . 1998. Insulin action in skeletal muscle from patients with NIDDM . Mol. Cell. Biochem . 182 :153–160. [PubMed] [Google Scholar]
- Zisman, A., Peroni O. D., Abel E. D., Michael M. D., Mauvais-Jarvis F., Lowell B. B., Wojtaszewski J. F., Hirshman M. F., Virkamaki A., Goodyear L. J., . et al. 2000. Targeted disruption of the glucose transporter 4 selectively in muscle causes insulin resistance and glucose intolerance . Nat. Med . 6 :924–928. doi: 10.1038/78693. [PubMed] [CrossRef] [Google Scholar]
- Zorzano, A., Muñoz P., Camps M., Mora C., Testar X., and Palacín M.. . 1996. Insulin-induced redistribution of GLUT4 glucose carriers in the muscle fiber. In search of GLUT4 trafficking pathways . Diabetes 45 Suppl 1 :S70–S81. doi: 10.2337/diab.45.1.s70. [PubMed] [CrossRef] [Google Scholar]
- Zumbaugh, M. D., Geiger A. E., Luo J., Shen Z., Shi H., and Gerrard D. E.. . 2021. O-GlcNAc transferase is required to maintain satellite cell function . Stem Cells 39 :945–958. doi: 10.1002/stem.3361. [PubMed] [CrossRef] [Google Scholar]
- Zurlo, F., Larson K., Bogardus C., and Ravussin E.. . 1990. Skeletal muscle metabolism is a major determinant of resting energy expenditure . J. Clin. Invest . 86 :1423–1427. doi: 10.1172/JCI114857. [PMC free article] [PubMed] [CrossRef] [Google Scholar]