Next Article in Journal
Thromboembolic Disease and Cardiac Thrombotic Complication in COVID-19: A Systematic Review
Previous Article in Journal
Integrative Analysis of the Metabolome and Transcriptome Provides Insights into the Mechanisms of Flavonoid Biosynthesis in Quinoa Seeds at Different Developmental Stages
Previous Article in Special Issue
Abscisic Acid Improves Insulin Action on Glycemia in Insulin-Deficient Mouse Models of Type 1 Diabetes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Protein O-GlcNAcylation in Metabolic Modulation of Skeletal Muscle: A Bright but Long Way to Go

College of Animal Science and Veterinary Medicine, Heilongjiang Bayi Agricultural University, Daqing 163319, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Metabolites 2022, 12(10), 888; https://doi.org/10.3390/metabo12100888
Submission received: 19 August 2022 / Revised: 9 September 2022 / Accepted: 17 September 2022 / Published: 22 September 2022
(This article belongs to the Special Issue New Pathways to Improve Muscle Metabolism and Muscle Growth)

Abstract

:
O-GlcNAcylation is an atypical, dynamic and reversible O-glycosylation that is critical and abundant in metazoan. O-GlcNAcylation coordinates and receives various signaling inputs such as nutrients and stresses, thus spatiotemporally regulating the activity, stability, localization and interaction of target proteins to participate in cellular physiological functions. Our review discusses in depth the involvement of O-GlcNAcylation in the precise regulation of skeletal muscle metabolism, such as glucose homeostasis, insulin sensitivity, tricarboxylic acid cycle and mitochondrial biogenesis. The complex interaction and precise modulation of O-GlcNAcylation in these nutritional pathways of skeletal muscle also provide emerging mechanical information on how nutrients affect health, exercise and disease. Meanwhile, we explored the potential role of O-GlcNAcylation in skeletal muscle pathology and focused on its benefits in maintaining proteostasis under atrophy. In general, these understandings of O-GlcNAcylation are conducive to providing new insights into skeletal muscle (patho) physiology.

1. Introduction

Various post-translational modifications (PTMs) continuously and dynamically amplify the proteome to produce a wide variety of protein forms and their biological functions, thereby harmonizing the proteome with the needs of the organism [1,2,3]. Glycosylation is one of the most widespread and abundant PTMs in living organisms, encompassing different types involving complex metabolic networks [4,5]. O-GlcNAcylation is a dynamic, reversible and atypical O-glycosylation that involves only the binding of a single N-acetylglucosamine (GlcNAc) to the serine and threonine residues of the target protein via a β-configuration O-glycosidic bond [6]. O-GlcNAc transferase (OGT) and O-GlcNAcase (OGA) are the only pair of mutually antagonistic process enzymes for protein O-GlcNAcylation [7]. OGT transfers GlcNAc from UDP-N-acetylglucosamine (UDP-GlcNAc) to the hydroxyl groups in the threonine and serine residues of the target proteins [8]. In contrast, OGA hydrolyzes GlcNAc from the O-GlcNAcylated proteins [9]. See Figure 1 for a schematic diagram of the O-GlcNAcylation process. Coordination of OGT and OGA accurately and rapidly modulate the O-GlcNAcylation cycle of thousands of proteins [10]. It is the biological properties of OGT and OGA that allow the rapid addition and removal of reversible GlcNAc multiple times during the life cycle of the target proteins that confer the high kinetics of O-GlcNAcylation [11,12]. UDP-GlcNAc is the only donor substrate providing the required GlcNAc for O-GlcNAcylation [13]. UDP-GlcNAc is produced from the hexosamine biosynthetic pathway (HBP) controlled by the input of glucose, glutamine, acetyl-CoA and uridine triphosphate (UTP), making it a sensor molecule for fluctuations in these macromolecules [14]. It is the physiological property of UDP-GlcNAc that endows O-GlcNAcylation with a unique high sensitivity to nutrient availability and thus serves as a critical nutrient sensor for the metabolism of carbohydrate, amino acid, lipid and nucleotide [15]. Meanwhile, O-GlcNAcylation senses external environmental disturbance and internal adverse stimulation, acts as a stress receptor integrating signaling pathway inputs from different partners, and generates spatiotemporal-specific adaptive molecules and physiological responses by targeting multiple substrates [16,17].
O-GlcNAcylation exists in almost all living organisms, and it is distributed in almost all human tissues, even in saliva and urine [18]. O-GlcNAcylation occurs in almost all cellular compartments, and its first discovery in 1984 overturned the conventional knowledge that glycosylation occurs only in the endoplasmic reticulum and Golgi apparatus [19]. It also shows that almost all functions of proteins in regulating various cellular processes are covered. This importance is reflected again by the of OGA and OGA defects lethality [20,21]. O-GlcNAcylation affects the function, activity, stabilization, localization and chaperone interactions of target proteins, thereby participating in the regulation of a series of cellular biological processes such as immunity [22,23], inflammation [24], autophagy [25,26], apoptosis [27], stemness [28,29], transcription [30], translation [31], signal transduction [32], mitochondrial function [33], epigenetics [10,34], chromatin remodeling [35], metabolic reprogramming [36] and cellular stress responses [37,38]. Meanwhile, O-GlcNAcylation has surprisingly extensive crosstalk with other PTMs such as phosphorylation, acetylation, ubiquitination, methylation, and so on [39]. In particular, the abundance and cycle time scales of O-GlcNAcylation are very similar to those of phosphorylation and thus form a yin–yang relationship that may be co-occurring or competing and negatively or positively affect each other through their interaction [40,41]. In addition, O-GlcNAcylation homeostasis is essential for the maintenance of the normal physiological function of cells, tissues and organs. Abnormal O-GlcNAcylation levels are closely associated with the pathogenesis and progression of various diseases such as neurodegenerative diseases [42], cardiovascular diseases [43,44], aging [45], obesity [46,47], amyotrophy [48], cancers [49], diabetes and diabetic complications [50].
To date, thousands of O-GlcNAcylated proteins have been identified in skeletal muscle, including contractile proteins, structural proteins, cytoskeletal proteins and sarcolemma protein as well as metabolic enzymes, transcription factors and signaling proteins and mitochondrial proteins [51,52,53,54,55]. This suggests a unique and central physiological role for O-GlcNAcylation in skeletal muscle. Recent data also suggest that O-GlcNAcylation is involved in different cellular processes in skeletal muscle such as regulation of contractile and structural properties, maintenance of energy metabolic homeostasis, and mediation of insulin resistance [48,56,57]. Meanwhile, O-GlcNAcylation plays a potential role in many diseases related to skeletal muscle defects such as neuromuscular diseases and amyotrophy [58,59]. What we are interested in is that fast and slow skeletal muscles have different fine characteristics of O-GlcNAcylation during rest, exercise or muscle atrophy [60]. However, these effects and features of O-GlcNAcylation are neglected or underestimated. This review highlights the spatiotemporal regulation of O-GlcNAcylation and promotes metabolic flexibility in skeletal muscle to maintain energy metabolic homeostasis. Meanwhile, this review emphasizes some consensus on the effects of exercise frequency, intensity, type and duration on O-GlcNAcylation and its process enzymes. In addition, this review discusses the potential role of O-GlcNAcylation in skeletal muscle-related disorders. These understandings of O-GlcNAcylation contribute to provide emerging insights into the (patho)physiology of skeletal muscle.

2. O-GlcNAcylation Is an Essential Metabolic Modulator in Skeletal Muscle Physiology

Skeletal muscle fibers are very abundant and metabolically active, particularly during exercise [61]. Skeletal muscle accounts for approximately 40% of the total body weight of mammals and contains up to 75% of total body proteins and 30% of the resting metabolic rate in adults [62]. Skeletal muscle is the main contributor of metabolic changes caused by exercise. Maximum exercise increases the whole-body metabolic rate 20 times higher than the resting value, while the adenosine triphosphate (ATP) conversion rate in working skeletal muscle is more than 100 times higher than that at rest [63]. The metabolic flexibility of skeletal muscle allows it to increase energy supply and provide sufficient “fuel” for contraction [64,65]. The importance of carbohydrate as a fuel source in exercise has been known, and the main source of these carbohydrate is glycogen reserve of skeletal muscle during exercise [66,67]. In addition, skeletal muscle actively secretes myogenic protein to exert autocrine, paracrine or endocrine functions to convey its energy demand to other organs, including adipose tissue, the liver, pancreas, cardiovascular system, brain, bone and skin, thus causing crosstalk [68,69]. O-GlcNAcylation plays a decisive role in skeletal muscle glucose homeostasis as a cellular trophic sensor by participating in glycolysis [70,71], tricarboxylic acid (TCA) cycle [36], insulin signaling [72,73] and glycogen metabolism [74,75].

2.1. O-GlcNAcylation Is an Extremely Sensitive Sensor for the Nutrition Availability in Skeletal Muscle via the HBP

It has been a popular view for many years that the cellular O-GlcNAcylation cycle is strictly controlled by HBP flow and UDP-GlcNAc concentration [76]. The HBP is a conventional branch of the glucose metabolic pathway and its flow varies with extracellular glucose level. See Figure 2 for the detailed process of HBP [77]. Initially, extracellular glucose is transported into the intracellular through glucose transporter 1–4 (GLUT1-4) [78]. Therefore, the O-GlcNAcylation cycle is strictly controlled by the flow of glucose through the HBP. Subsequently, intracellular glucose was phosphorylated to glucose-6-phosphate by hexokinase, and then glucose-6-phosphate was further isomerized to fructose-6-phosphate by phosphoglucose isomerase [79]. Three to 5% of fructose-6-phosphate was added with an amino group from a glutamine to synthesize glucosamine-6-phosphate and a glutamate by glutamine fructose-6-phosphate amidotransferase-1 [80]. This enzymatic reaction is the rate-limiting step of HBP, and glutamine fructose-6-phosphate amidotransferase-1 is also the critical rate-limiting enzyme of HBP. Glutamine is necessary for this enzymatic reaction, but this restriction is bypassed by glucosamine as an extended supplement [81]. Glucosamine-6-phosphate acetyltransferase acetylates glucosamine-6-phosphate to N-acetylglucosamine-6-phosphate using acetyl-CoA [82]. Then, N-acetylglucosamine-6-phosphate is catalytically translocated to N-acetylglucosamine-1-phosphate by N-acetylglucosamine-phosphate mutase 1. Finally, UTP is then utilized by UDP-N-acetylglucosamine pyrophosphorylase-1 to convert N-acetylglucosamine-1-phosphate into UDP-GlcNAc and release Pi [83]. UDP-GlcNAc, as the end product of HBP, involves the participation of glucose, glutamine, uridine, acetyl-CoA and ATP [84]. Therefore, UDP-GlcNAc is sensitive to the fluctuation of all these nutrients as a “sensor molecule” [85]. The GlcNAc provided by UDP-GlcNAc is necessary and irreplaceable for O-GlcNAcylation. It is this physiological property that makes O-GlcNAcylation a major metabolic node that integrates the metabolism of carbohydrates, amino acids, lipids and nucleotides [86]. On this basis, combined with the high dynamic nature of O-GlcNAcylation, it makes it a highly sensitive and rapidly responsive metabolic sensor of nutrient availability reporting multiple pathway functional states [87]. Overall, HBP and UDP-GlcNAc link altered metabolisms and O-GlcNAcylation, providing an important mechanism for cells to perceive and respond to nutritional availability.

2.2. O-GlcNAcylation Is the Pivotal Modulator of Glucose Metabolic Homeostasis in Skeletal Muscle

During low-to-moderate-intensity exercise, the main fuel sources supplying skeletal muscle are glucose and are heavily influenced by insulin and glucose utilization increases progressively with increasing exercise intensity until near maximum intensity [64]. Several recent studies have demonstrated that O-GlcNAcylation controls early glucose metabolism in skeletal muscle [48,72,88,89]. The fuel metabolism and energy homeostasis of skeletal muscle are greatly dependent on the glucose uptake and disposal. O-GlcNAcylation alters GLUT4 translocation, prevents GLUT4 phosphorylation, controls GLUT4 downstream signal transduction, or directly controls vesicle proteins to sustain glucose absorption [90,91]. Meanwhile, O-GlcNAcylation of hypoxia-inducible factor-1 enhances GLUT1 transcription and thereby regulates glucose uptake [92]. Additionally, O-GlcNAcylated hexokinase IV or glucokinase upregulates their expression, which is of positive significance in controlling glucose flow [93]. Similar alterations have taken place in O-GlcNAcylated phosphoglucose isomerase [88]. O-GlcNAcylation of glucose-6-phosphate dehydrogenase at Ser84 enhances its catalytic activity, thereby forcing a shift in glucose flow to the pentose phosphate pathway [94]. The O-GlcNAcylation of phosphofructokinase 1 at Ser529 inhibits its activity and redirects glucose flow into the pentose phosphate pathway, which also reduces glycolytic glucose flow [71]. The O-GlcNAcylation 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase at Ser172 competes with its phosphorylation to modulate glycolysis [95]. In fact, almost all enzymes involved in the glycolysis pathway are O-GlcNAcylated such as fructose-1,6-bisphosphate aldolase and glyceraldehyde-3-phosphate dehydrogenase [7]. Transcription of these enzymes is also mediated by O-GlcNAcylation of Sp1 [96]. Another critical metabolic enzyme in the glycolytic pathway, phosphoglycerate kinase 1, is also O-GlcNAcylated at Thr255, which enhances its activity to promote glycolysis and translocation to mitochondria to inhibit the TCA cycle [36]. The downstream phosphoglycerate mutase and α-enolase were also identified as O-GlcNAcylated, but their functions need to be further investigated. The O-GlcNAcylation of pyruvate kinase M2 at Thr405 and Ser406 destroys its stability, reduces its activity and causes its nuclear translocation [70]. These findings imply that O-GlcNAcylation regulates the skeletal muscle’s glycolysis process fundamentally [97]. In addition, the terminal lactate dehydrogenase is also O-GlcNAcylated, indicating that O-GlcNAcylation modulates the utilization of glycolysis end products [88]. Skeletal muscle is the largest glycogen reservoir and has four times the capacity of the liver [98]. Glycogen synthesis and catabolism contribute to the maintenance of energy homeostasis in skeletal muscle, and O-GlcNAcylation is a regulator of glycogen metabolism. For example, the O-GlcNAcylation of glycogen synthase kinase 3β (GSK3β) competes with its phosphorylation at Ser9 to inhibit its activity [75,99]. Similar regulation also occurs on the O-GlcNAcylated glycogen synthase (GS) [100]. In addition, OGT indirectly regulates the global metabolism of skeletal muscle through mediated expression of interleukin-6 and interleukin-15 [57,101]. Skeletal muscle is also one of the main sites of energy-producing lipid metabolism. Free fatty acids released by lipolysis in the liver and adipose tissue have a significant contribution to the substrate supply of muscle contraction during exercise. The increased HBP flux and O-GlcNAcylation levels are associated with greater fatty acid oxidation in the heart, possibly through O-GlcNAcylation of CD36 (the fatty acid transporter) [102]. Similar results were seen in adipocytes and short OGA, a splice variant of OGA, is associated with lipid droplets. In addition, fatty acids and their derivatives may play different regulatory roles on the quality and function of skeletal muscle [61,65]. The O-GlcNAcylation of carbohydrate-responsive element-binding protein, farnesoid X receptor, sterol regulatory element binding protein 1, perilipin 1, liver X receptor-α, fatty acid synthase and other proteins is also involved in the regulation of lipid metabolism, but this information comes from the liver and adipose tissue, etc., rather than skeletal muscle [46,103,104,105,106,107]. However, the connection of information and energy between skeletal muscle and other organs or tissues is close and frequent [69]. Therefore, this information is of certain value for understanding the modulation of O-GlcNAcylation on skeletal muscle energy homeostasis, but we do not discuss it further.

2.3. O-GlcNAcylation Is the Precise Spatiotemporal Regulator of Insulin Signal Transduction in Skeletal Muscle

Skeletal muscle is the main tissue for glucose treatment under insulin-stimulated conditions and thus plays a pivotal role in blood glucose control and systemic metabolic homeostasis [108]. It has long been generally recognized that there is a strong correlation between the excessive HBP flow and UDP-GlcNAc level and skeletal muscle insulin resistance and diabetic pathology [109,110,111]. The concentration of UDP-GlcNAc and HBP flow have a significant impact on OGT’s catalytic activity [112]. Meanwhile, co-infusion of insulin and glucosamine resulted in increased HBP flux and UDP-GlcNAc levels enhancing overall O-GlcNAcylation levels and many O-GlcNAcylated proteins on skeletal muscle [113,114]. Interestingly, almost all of the major effector molecules of the insulin signaling pathway are O-GlcNAcylated, such as insulin receptor substrates-1 (IRS-1), phosphoinositide 3-kinase (PI3K), 3-phosphoinositide-dependent kinase-1 (PDK-1), protein kinase B (AKT), and forkhead box O1 (FoxO1). Therefore, it is self-evident that this chronically elevated O-GlcNAcylation disrupts the insulin signal transduction in skeletal muscle [115]. That is, the mechanism of HBP-induced skeletal muscle insulin resistance is that the increased O-GlcNAcylation of participants in insulin signaling disrupts the balance of its antagonism with phosphorylation on these proteins to negatively regulate insulin signaling [40,116]. See Figure 3 for more details. Insulin receptor binds to insulin to make itself phosphorylated on the cytomembrane under normal physiological conditions, which cause subsequent signal cascade [117]. Subsequently, IRS-1 is phosphorylated, resulting in PI3K activation [118]. Then, activated PI3K produces phosphatidylinositol-3,4,5-trisphosphate (PIP3) on the cytomembrane, which binds to AKT and PDK1 [119,120]. AKT is phosphorylated at the Thr308 by PDK1 to activate [121]. Downstream target proteins are phosphorylated by activated AKT for activation or inhibition [122]. Phosphorylation inhibition of downstream GSK3β and GS promotes glycogen synthesis [74]. FoxO1 phosphorylation inhibits gluconeogenesis gene transcription, thereby reducing gluconeogenesis [123]. Membrane translocation of GLUT4 is promoted to glycolysis and glucose uptake [124]. Ultimately, these pathways make the elevated glucose return to homeostasis [125]. However, PIP3 also recruits OGT to the cytomembrane through the strong interaction with the phosphoinositide-binding domain of OGT under diabetes or another insulin-insensitive state such as the prolonged postprandial response [126]. Meanwhile, OGT is phosphorylated and enhances its activity through IR [127]. Subsequently, OGT catalyzes the dynamic O-GlcNAcylation of IRS, PDK1, AKT, FoxO1 and other insulin signal molecules [128,129]. The O-GlcNAcylation of IRS-1 at Ser1101 and IRS-2 at Ser1149 inhibits their phosphorylation at the same site, resulting in the attenuation of insulin signal [128]. The O-GlcNAcylation of AKT at the Thr305 and Thr312 inhibits its phosphorylation at Thr308 through disrupting the interaction between AKT and PDK1 [130]. The O-GlcNAcylations of PI3K and PDK1 are also implicated in insulin-signaling attenuation. In fact, the precise control of insulin-signaling players by dynamic antagonism of phosphorylation and O-GlcNAcylation goes far beyond this. For example, O-GlcNAcylation of protein-tyrosine phosphatase 1B regulates its phosphatase activity and affects insulin signaling, thus causing insulin resistance [131,132]. O-GlcNAcylation of S6 kinase beta1 inhibits its phosphorylation and mTORC1 signaling, thereby causing insulin resistance [133]. In addition, O-GlcNAcylation of cAMP-response element binding protein (CREB) crosstalk with its phosphorylation [134,135]. O-GlcNAcylation of CREB-regulated transcription coactivator 2 at Ser70 and Ser171 inhibits the phosphorylation of these two sites, thereby facilitating their nuclear translocation and binding to CREB [136].

2.4. O-GlcNAcylation Is the Pivotal Maintainer of the TCA Cycle and Mitochondrial Homeostasis in Skeletal Muscle

Skeletal muscle is rich in mitochondria and relies heavily on oxidative phosphorylation to generate energy [137]. During vigorous exercise, intramuscular oxygen consumption and local blood flow increased significantly by more than 30-fold and the estimated TCA cycle flux increased by 70- to 100-fold [138]. Many studies have shown that changes in cellular O-GlcNAcylation levels alter mitochondrial function, especially under stress [139,140]. Mitochondrial OGT, one of the OGT splice variants, has a unique mitochondrial targeting sequence [141]. Although these data have shown that mitochondrial proteins are targets of O-GlcNAcylation, O-GlcNAcylation studies were initially hampered by the unclear mechanism of UDP-GlcNAc translocation to mitochondria. It was once thought that OGT in mitochondria was rare or even non-existent. This obstacle was overcome by Gerald W. Hart in 2015, whose study found that the mitochondrial pyrimidine nucleotide carrier 1 is responsible for efficient transport of radioactive 3 H UDP-GlcNAc into mitochondria [45]. This study also demonstrates the presence of mitochondria-specific OGA, which also proves the existence of an active O-GlcNAcylation cycle. This O-GlcNAcylation biology’s founder has once again advanced the understanding of O-GlcNAcylation in mitochondria. Subsequently, O-GlcNAcylation on many mitochondrial respiratory chain complexes was identified such as complex I, II, and IV, which led to a decrease in their activity and mitochondrial calcium and cellular ATP content [142]. Meanwhile, eleven O-GlcNAcylated proteins were identified in mitochondria, all of which are metabolic enzymes involved in the TCA cycle and ATP synthesis [143]. It should be emphasized that almost all metabolic enzymes involved in the TCA cycle on mitochondrion are also O-GlcNAcylated such as pyruvate dehydrogenase, aconitase, isocitrate dehydrogenase, ɑ-ketoglutarate dehydrogenase, succinyl-CoA ligase, succinate dehydrogenase and malate dehydrogenase, except for citrate synthase and fumarase [88,144]. Further, two proteomic studies identified up to hundreds of mitochondrial proteins as targets of O-GlcNAcylation, most of which are involved in a variety of biological processes such as oxidative phosphorylation, urea cycle, fatty acid oxidation and calcium regulation [145,146,147]. These data emphasize the importance of O-GlcNAcylation as a regulatory molecule of mitochondrial metabolism and respiration. Meanwhile, O-GlcNAcylation modifies a variety of mitochondrial proteins to change the motility, morphology and quality of mitochondria [148]. For example, O-GlcNAcylation of peroxisome proliferators-activated receptor γ coactivator ɑ is beneficial to maintain mitochondrial biogenesis [149]. O-GlcNAcylation of Milton in response to the glucose availability mediates changes in neuronal mitochondrial mobility [150]. Meanwhile, O-GlcNAcylation of dynamin-related protein 1 at Thr585 and Thr586 inhibits its phosphorylation at Ser637, which promoted its translocation to mitochondria and increased fragmentation as well as decreased membrane potential [151,152,153]. These data have demonstrated that O-GlcNAcylation is closely related to mitochondrial function. On the other hand, O-GlcNAcylation should play an important role in mitochondrial oxidative stress. For example, Increased O-GlcNAcylation levels mediated by glucosamine treatment and OGT overexpression attenuated hydrogen peroxide-induced mitochondrial membrane potential and enhanced the recruitment of the anti-apoptotic protein B-cell lymphoma-2 (Bcl-2) to mitochondria [154]. In addition, a subsequent study demonstrated that Bcl-2 is O-GlcNAcylated [155]. O-GlcNAcylation of 8-oxoguanine DNA glycosylase impairs mitochondrial DNA damage repair under oxidative stress [156]. O-GlcNAcylation of voltage-dependent anion channel 1 attenuates oxidative stress-induced loss of mitochondrial membrane potential [157]. O-GlcNAcylation of superoxide dismutase 2 has a similar effect [158].

3. The Fine Characteristics of O-GlcNAcylation, Skeletal Muscle Fibre Types and Exercises

Skeletal muscle is a heterogeneous cell population consisting of different types of muscle fibers that is classified as slow-twitch muscle fibers (type I) or fast-twitch muscle fibers (types IIa, IIx, and IIb) according to the contractile property of “time-to-peak tension” or “twitch” characteristics [159]. The heterogeneity allows the same muscle to be used for a variety of tasks, from continuous low-intensity activities (such as postures), to repetitive submaximal contractions (such as exercise), to fast and intense maximal contractions (such as jump, kick) [160]. Each muscle fiber type has its own specific and unique contribution to respond to functional needs in the best way [161]. Type I is heavily dependent on oxidative metabolism for the continuous work, and Type I has a red appearance due to its rich myoglobin and oxidizing enzymes [162]. Type II is characterized by glycolytic metabolism for the phased work, but types IIa, IIx and IIb have different degrees of oxidation and glycolytic capacity. This also leads to the intermediate appearance of type IIa and the white appearance of type IIx and IIb [163]. Meanwhile, type IIb is mainly exist in rodents, and types IIa and IIx are classified in humans [164]. It is the metabolic and morphological phenotype of each skeletal muscle fiber type coupled with the essential regulation of O-GlcNAcylation on skeletal muscle metabolic homeostasis and structural and contractile properties leading to the unique fine characteristics of O-GlcNAcylation in each type of skeletal muscle fiber [48,72,144,165,166]. In fact, the modulation of O-GlcNAcylation, OGT and OGA in each muscle fiber type is completely different, which may be attributed to the patterns of metabolic and molecular responses unique to each skeletal muscle fiber type [167]. For example, the global O-GlcNAcylation was much higher in slow-twitch muscle at rest [72,168]. Meanwhile, the O-GlcNAcylation process enzyme and glutamine fructose-6-phosphate amidotransferase in soleus were higher than those in the extensor digitorum longus muscle, and so were their activities [168,169]. The O-GlcNAcylation pattern of each muscle fiber type is also differentiated in health and in diseases such as atrophy and aging [170,171,172]. For example, the overall level of O-GlcNAcylation in soleus was decreased under the disuse atrophy model. The activities of OGT and OGA in atrophic soleus muscle increased and decreased respectively at the later stage of this model [169]. Changes of O-GlcNAcylation in each muscle fiber type during exercise are more complex and unpredictable. The global O-GlcNAcylation in both slow- and fast-twitch muscle levels were unchanged during a single acute exercise, and its fine regulation was only observed on some specific myofilament proteins such as myosin light chain 2 [168,173]. This fine regulation was reversed after recovery. Confusingly, another study showed that the same acute exercise increased the overall O-GlcNAcylation, regardless of slow- or fast-twitch muscle [174]. This confusion is attributed to the differentiation of O-GlcNAcylation on the metabolic flexibility and glucose utilization and oxidative stress of skeletal muscle in the short term [64,175,176]. Chronic exercise involves the adaptation and remodeling of skeletal muscle [60]. Due to skeletal muscle plasticity, muscle fibers remodel their structural and functional properties and protein homeostasis in response to the different environmental conditions imposed on them, such as nerve stimulation pattern, load intensity, substrate availability and hormone signal changes [177,178,179,180]. Meanwhile, the changes of mechanical strain, calcium flux, ATP turnover, redox balance, ROS production and oxygen pressure caused by contraction are all related to the activation of signal transduction cascades that regulate skeletal muscle plasticity [181,182]. A series of changes in myosin isoforms, protein turnover, metabolism, mitochondrial function, intracellular signaling or transcription occur during skeletal muscle remodeling [183]. These processes also have the close participation of O-GlcNAcylation [184,185,186,187,188]. The muscle fiber type is genetically determined during development, and the adaptive “transformation” of muscle fiber type from one to another is still hotly debated [189]. The regulatory effect of O-GlcNAcylation on differentiation and myogenesis have been demonstrated [89,190,191,192]. For example, the global O-GlcNAcylation reduced in early myogenesis before myoblast fusion and the inactivation of OGA severely interferes with the expression of myogenin and myogenic regulatory factor 4, which indicate that the terminal differentiation program of skeletal myogenesis is negatively regulated by O-GlcNAcylation [193]. Hyper O-GlcNAcylation of myocyte-specific enhancer factor 2c at Thr9 inhibits its DNA binding affinity to the myogenin promoter to attenuate skeletal muscle differentiation [194]. This negative regulation was also eliminated by hypo O-GlcNAcylation of myocyte-specific enhancer factor 2d [195]. Skeletal muscle regeneration and repair can provide functional recovery during chronic exercise [196]. O-GlcNAcylation contributes to a self-renewal ability of muscle satellite cells after acute injury and plays a pivotal role in ensuring the health and function of skeletal muscle [196]. Although the changes of O-GlcNAcylation still largely depend on the skeletal muscle fiber type and exercise program, there is still a long way to go to reach a consensus. Another awkward situation is that the nodes of O-GlcNAcylation associated with exercise have been studied more in cardiac muscle and smooth muscle than in skeletal muscle [60,197]. We do not discuss further here.

4. O-GlcNAcylation, Autophagy and Skeletal Muscle Pathology

There is abundant evidence to support the potential role of O-GlcNAcylation in the skeletal muscle pathology [88,198]. For example, the global O-GlcNAcylation level in skeletal muscle is increased and its localization is abnormally altered in some human neuromuscular diseases such as muscular dystrophy, myositis and rhabdomyolysis [199]. Abnormal changes of the global O-GlcNAcylation level abnormal change under amyotrophic lateral sclerosis, and the progression of this disease is may be attributed to the O-GlcNAcylation of the tail domain of neurofilament protein M in spinal cord tissue [200,201]. The O-GlcNAcylation of gigaxonin at Ser272 and Thr277 mediates intermediate filament proteostasis and turnover to regulate human giant axonal neuropathy [202]. UDP-N-acetylglucosamine 2-epimerase/N-acetylmannosamine kinase (GNE) is a bifunctional enzyme critical for sialic acid biosynthesis [203]. GNE mutations cause GNE myopathy such as hereditary inclusion body myopathy and distal myopathy with rimmed vacuoles, and O-GlcNAcylation of GNE at Thr743 cooperates with the phosphorylation at the same site to affect its efficiency and activity of GNE [204]. However, the abnormal overall O-GlcNAcylation levels is the major pathogenic factors in skeletal muscle atrophy [48,205]. For example, disruption of OGA activity mediates hyper O-GlcNAcylation, leading to muscle atrophy [58]. The possible underlying mechanisms of O-GlcNAcylation with regard to skeletal muscle atrophy are the following pathways: (1) as we previously discussed, O-GlcNAcylation seems to play a negative regulatory role on myogenesis as well as differentiation; (2) as we previously discussed, the differential regulation of O-GlcNAcylation on metabolism, signal transduction and stress response; (3) the regulation of O-GlcNAcylation on skeletal muscle contraction and structural properties such as calcium affinity, which is the most discussed topic at present. For details, please read to reference [56]; (4) the regulation of O-GlcNAcylation on synthesis and decomposition such as preventing proteasome degradation and mediating autophagy process and maintaining proteostasis [206], and this is what we need to discuss below. Skeletal muscle constantly finely regulates the movement and posture of the body, especially in daily exercise and physical activity, which produce a large amount of ROS to disrupt the normal physiological functions [207,208]. Therefore, skeletal muscle must not only efficiently recycle damaged or aged organelles and accumulated protein aggregates, but also efficiently decompose proteins to meet the energy requirements [209]. Autophagy, as a nutritional mechanism, is essential for removing aging and damaged cellular components, decomposing unallocated nutrient reserves and remodeling cellular structure [210]. The critical role of proper autophagic flux for skeletal muscle function and metabolism cannot be overstated [211]. Indeed, autophagy has been considered as an emerging metabolic regulatory pathway in skeletal muscle, and its role in skeletal muscle metabolism and atrophic diseases has been extensively studied [212,213]. In skeletal muscle autophagy signal transduction, mTOR, AMPK, AKT and FoxO3, as important central trophic effectors, regulate autophagy process according to nutritional status, and their interaction with OGT is inseparable [214,215,216,217]. Given the critical role of O-GlcNAcylation in these processes, O-GlcNAcylation may affect autophagy flux and proteostasis to regulate skeletal muscle mass and proteolysis [206]. For example, O-GlcNAcylation regulates autophagy flux and proteasome activity in the face of protein toxicity challenges [218]. A similar role in huntingtin toxicity was observed [219]. It should be noted that O-GlcNAcylation affects proteasome function and activity as well as assembly and changes cellular protein homeostasis [220]. O-GlcNAcylation of Rpt2 inhibits its activity to shut down the proteasome and cause protein accumulation [221,222]. O-GlcNAcylation of ryanodine receptor 1 increased ubiquitination and proteasomal degradation, thus triggering the cleavage of desmin filaments and the destruction of myofibrils [223]. Meanwhile, Caenorhabditis elegans bearing null mutations of ogt-1 or oga-1 increases autophagy under food deprivation-induced starvation [224,225]. Genetic ablation of OGT in mouse livers also affects autophagic flux under similar deprivation [25,216]. For example, the reduction of OGT in Drosophila was associated with the increase of autophagy-related gene (ATG) product and autophagy lysosomes [226]. OGT and O-GlcNAcylation in Caenorhabditis elegans affect the fusion of autophagosomes and lysosomes [227]. The O-GlcNAcylation level of synaptosome-associated protein 29 is also affected by nutritional availability, which eliminates its involvement in soluble N-ethylmaleimide-sensitive factor attachment protein receptor complex formation and thus disrupts the fusion between autophagosomes and endosomes/lysosomes [26,228]. Similarly, O-GlcNAcylation of Golgi reassembly stacking protein 55 acts as a tether to promote autophagosome maturation [229]. In addition, O-GlcNAcylation also participate in autophagy regulation by modifying other pivotal autophagy regulators such as beclin1, Bcl-2 and UNC-51-like kinase 1 (ULK1) [230]. See Figure 4 for more details.

5. Conclusions and Perspectives

In the last four decades, O-GlcNAcylation in skeletal muscle physiology has received increasing attention, but these understandings about its complex regulation and molecular mechanisms are still inadequate and initial. O-GlcNAcylation is a potential essential regulator for skeletal muscle glucose homeostasis and insulin sensitivity and mitochondrial biogenesis. Global O-GlcNAcylation patterns in skeletal muscle vary with rest, exertion or exercise and fiber type and thus differentially regulate a series of the pivotal signaling pathways and cellular stress responses as well as skeletal muscle plasticity and adaptation. A deeper understanding of these precise mechanisms by which O-GlcNAcylation functions is well worthwhile and a future hotspot. This future direction may lead to new valuable insights into skeletal muscle metabolism and exercise. O-GlcNAcylation is considered as a potential emerging mechanism to regulate contractile and structural properties of skeletal muscle. Future relevant studies should be considered in identifying skeletal muscle O-GlcNAcylated proteins critical for contractile activity and sarcomere cytoskeleton and their precise sites, while further focusing on their interactions and the dynamic crosstalk of O-GlcNAcylation with phosphorylation and other PTMs. The balance of phosphorylation and O-GlcNAcylation on myosin light chain 2 is a classic excellent example. O-GlcNAcylation is involved in skeletal muscle pathology in some related diseases such as atrophy, neuromuscular disease and insulin resistance, and it remains to be discussed as the cause or consequence of skeletal muscle impairments. O-GlcNAcylation has been recognized for mediating proteostasis in skeletal muscle through its benefits in autophagy, apoptosis and proteasomes. It deserves constant attention and the regulation of O-GlcNAcylation on myogenesis, differentiation and signaling pathways that maintain synthesis/degradation balance in skeletal muscle should not be neglected. Over time, further investigation of the pathological role of O-GlcNAcylation in skeletal muscle will provide the multiple novel therapeutic targets and the basis for developing new strategies to combat certain skeletal muscle diseases or metabolic disorders. There is a long way to go, although the future is bright and ambitious.

Author Contributions

Y.L. conceptualized and wrote the manuscript. Y.H. performed a literature search and made these pictures. S.L. strictly supervised the whole work and revised the manuscript. The final manuscript was read and approved by all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (31972637), the Key Program of the Natural Science Foundation of Heilongjiang Province (ZD2019C004) and the Graduate Innovative Research Project of Heilongjiang Bayi Agricultural University (YJSCX2021-Z01).

Acknowledgments

Thanks to Yuying Yang, Bin Xu, Hongming Lv and Shuai Lian for their guidance provided in the writing process.

Conflicts of Interest

There is no conflict of interest in the final version.

References

  1. Conibear, A.C. Deciphering protein post-translational modifications using chemical biology tools. Nat. Rev. Chem. 2020, 4, 674–695. [Google Scholar] [CrossRef]
  2. Altelaar, A.F.; Munoz, J.; Heck, A.J. Next-generation proteomics: Towards an integrative view of proteome dynamics. Nat. Rev. Genet. 2013, 14, 35–48. [Google Scholar] [CrossRef]
  3. Aebersold, R.; Agar, J.N.; Amster, I.J.; Baker, M.S.; Bertozzi, C.R.; Boja, E.S.; Costello, C.E.; Cravatt, B.F.; Fenselau, C.; Garcia, B.A.; et al. How many human proteoforms are there? Nat. Chem. Biol. 2018, 14, 206–214. [Google Scholar] [CrossRef] [PubMed]
  4. Schjoldager, K.T.; Narimatsu, Y.; Joshi, H.J.; Clausen, H. Global view of human protein glycosylation pathways and functions. Nat. Rev. Mol. Cell Biol. 2020, 21, 729–749. [Google Scholar] [CrossRef] [PubMed]
  5. Reily, C.; Stewart, T.J.; Renfrow, M.B.; Novak, J. Glycosylation in health and disease. Nat. Rev. Nephrol. 2019, 15, 346–366. [Google Scholar] [CrossRef]
  6. Saha, A.; Bello, D.; Fernández-Tejada, A. Advances in chemical probing of protein O-GlcNAc glycosylation: Structural role and molecular mechanisms. Chem. Soc. Rev. 2021, 50, 10451–10485. [Google Scholar] [CrossRef]
  7. Chatham, J.C.; Zhang, J.; Wende, A.R. Role of O-Linked N-Acetylglucosamine Protein Modification in Cellular (Patho)Physiology. Physiol. Rev. 2021, 101, 427–493. [Google Scholar] [CrossRef]
  8. Joiner, C.M.; Levine, Z.G.; Aonbangkhen, C.; Woo, C.M.; Walker, S. Aspartate Residues Far from the Active Site Drive O-GlcNAc Transferase Substrate Selection. J. Am. Chem. Soc. 2019, 141, 12974–12978. [Google Scholar] [CrossRef]
  9. Li, B.; Li, H.; Hu, C.W.; Jiang, J. Structural insights into the substrate binding adaptability and specificity of human O-GlcNAcase. Nat. Commun. 2017, 8, 666. [Google Scholar] [CrossRef]
  10. Yang, X.; Qian, K. Protein O-GlcNAcylation: Emerging mechanisms and functions. Nat. Rev. Mol. Cell Biol. 2017, 18, 452–465. [Google Scholar] [CrossRef] [Green Version]
  11. Kohler, J.J. Carb cutting works better with a partner. Nat. Struct. Mol. Biol. 2017, 24, 433–435. [Google Scholar] [CrossRef] [PubMed]
  12. Pathak, S.; Alonso, J.; Schimpl, M.; Rafie, K.; Blair, D.E.; Borodkin, V.S.; Albarbarawi, O.; van Aalten, D.M.F. The active site of O-GlcNAc transferase imposes constraints on substrate sequence. Nat. Struct. Mol. Biol. 2015, 22, 744–750. [Google Scholar] [CrossRef] [PubMed]
  13. Levine, Z.G.; Walker, S. The Biochemistry of O-GlcNAc Transferase: Which Functions Make It Essential in Mammalian Cells? Annu. Rev. Biochem. 2016, 85, 631–657. [Google Scholar] [CrossRef] [PubMed]
  14. Ruan, H.B.; Singh, J.P.; Li, M.D.; Wu, J.; Yang, X. Cracking the O-GlcNAc code in metabolism. Trends Endocrinol. Metab. TEM 2013, 24, 301–309. [Google Scholar] [CrossRef]
  15. Fehl, C.; Hanover, J.A. Tools, tactics and objectives to interrogate cellular roles of O-GlcNAc in disease. Nat. Chem. Biol. 2022, 18, 8–17. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Y.; Yao, R.Z.; Lian, S.; Liu, P.; Hu, Y.J.; Shi, H.Z.; Lv, H.M.; Yang, Y.Y.; Xu, B.; Li, S.Z. O-GlcNAcylation: The “stress and nutrition receptor” in cell stress response. Cell Stress Chaperones 2021, 26, 297–309. [Google Scholar] [CrossRef]
  17. Han, C.; Gu, Y.; Shan, H.; Mi, W.; Sun, J.; Shi, M.; Zhang, X.; Lu, X.; Han, F.; Gong, Q.; et al. O-GlcNAcylation of SIRT1 enhances its deacetylase activity and promotes cytoprotection under stress. Nat. Commun. 2017, 8, 1491. [Google Scholar] [CrossRef]
  18. Wulff-Fuentes, E.; Berendt, R.R.; Massman, L.; Danner, L.; Malard, F.; Vora, J.; Kahsay, R.; Olivier-Van Stichelen, S. The human O-GlcNAcome database and meta-analysis. Sci Data 2021, 8, 25. [Google Scholar] [CrossRef]
  19. Xu, S.; Tong, M.; Suttapitugsakul, S.; Wu, R. Spatial and temporal proteomics reveals the distinct distributions and dynamics of O-GlcNAcylated proteins. Cell Rep. 2022, 39, 110946. [Google Scholar] [CrossRef]
  20. Muha, V.; Authier, F.; Szoke-Kovacs, Z.; Johnson, S.; Gallagher, J.; McNeilly, A.; McCrimmon, R.J.; Teboul, L.; van Aalten, D.M.F. Loss of O-GlcNAcase catalytic activity leads to defects in mouse embryogenesis. J. Biol. Chem. 2021, 296, 100439. [Google Scholar] [CrossRef]
  21. Moulton, M.J.; Humphreys, G.B.; Kim, A.; Letsou, A. O-GlcNAcylation Dampens Dpp/BMP Signaling to Ensure Proper Drosophila Embryonic Development. Dev. Cell 2020, 53, 330–343.e333. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, J.; Qian, C.; Cao, X. Post-Translational Modification Control of Innate Immunity. Immunity 2016, 45, 15–30. [Google Scholar] [CrossRef] [PubMed]
  23. Chang, Y.H.; Weng, C.L.; Lin, K.I. O-GlcNAcylation and its role in the immune system. J. Biomed. Sci. 2020, 27, 57. [Google Scholar] [CrossRef] [PubMed]
  24. Li, X.; Gong, W.; Wang, H.; Li, T.; Attri, K.S.; Lewis, R.E.; Kalil, A.C.; Bhinderwala, F.; Powers, R.; Yin, G.; et al. O-GlcNAc Transferase Suppresses Inflammation and Necroptosis by Targeting Receptor-Interacting Serine/Threonine-Protein Kinase 3. Immunity 2019, 50, 576–590.e576. [Google Scholar] [CrossRef] [PubMed]
  25. Ruan, H.B.; Ma, Y.; Torres, S.; Zhang, B.; Feriod, C.; Heck, R.M.; Qian, K.; Fu, M.; Li, X.; Nathanson, M.H.; et al. Calcium-dependent O-GlcNAc signaling drives liver autophagy in adaptation to starvation. Genes Dev. 2017, 31, 1655–1665. [Google Scholar] [CrossRef]
  26. Guo, B.; Liang, Q.; Li, L.; Hu, Z.; Wu, F.; Zhang, P.; Ma, Y.; Zhao, B.; Kovács, A.L.; Zhang, Z.; et al. O-GlcNAc-modification of SNAP-29 regulates autophagosome maturation. Nat. Cell Biol. 2014, 16, 1215–1226. [Google Scholar] [CrossRef]
  27. Zhang, B.; Li, M.D.; Yin, R.; Liu, Y.; Yang, Y.; Mitchell-Richards, K.A.; Nam, J.H.; Li, R.; Wang, L.; Iwakiri, Y.; et al. O-GlcNAc transferase suppresses necroptosis and liver fibrosis. JCI Insight 2019, 4, e127709. [Google Scholar] [CrossRef]
  28. Hao, Y.; Fan, X.; Shi, Y.; Zhang, C.; Sun, D.E.; Qin, K.; Qin, W.; Zhou, W.; Chen, X. Next-generation unnatural monosaccharides reveal that ESRRB O-GlcNAcylation regulates pluripotency of mouse embryonic stem cells. Nat. Commun. 2019, 10, 4065. [Google Scholar] [CrossRef]
  29. Jang, H.; Kim, T.W.; Yoon, S.; Choi, S.Y.; Kang, T.W.; Kim, S.Y.; Kwon, Y.W.; Cho, E.J.; Youn, H.D. O-GlcNAc regulates pluripotency and reprogramming by directly acting on core components of the pluripotency network. Cell Stem Cell 2012, 11, 62–74. [Google Scholar] [CrossRef]
  30. Hart, G.W. Nutrient regulation of signaling and transcription. J. Biol. Chem. 2019, 294, 2211–2231. [Google Scholar] [CrossRef] [Green Version]
  31. Li, X.; Zhu, Q.; Shi, X.; Cheng, Y.; Li, X.; Xu, H.; Duan, X.; Hsieh-Wilson, L.C.; Chu, J.; Pelletier, J.; et al. O-GlcNAcylation of core components of the translation initiation machinery regulates protein synthesis. Proc. Natl. Acad. Sci. USA 2019, 116, 7857–7866. [Google Scholar] [CrossRef] [PubMed]
  32. Wells, L.; Vosseller, K.; Hart, G.W. Glycosylation of nucleocytoplasmic proteins: Signal transduction and O-GlcNAc. Science 2001, 291, 2376–2378. [Google Scholar] [CrossRef] [PubMed]
  33. Trapannone, R.; Mariappa, D.; Ferenbach, A.T.; van Aalten, D.M. Nucleocytoplasmic human O-GlcNAc transferase is sufficient for O-GlcNAcylation of mitochondrial proteins. Biochem. J. 2016, 473, 1693–1702. [Google Scholar] [CrossRef] [PubMed]
  34. Forma, E.; Jóźwiak, P.; Bryś, M.; Krześlak, A. The potential role of O-GlcNAc modification in cancer epigenetics. Cell. Mol. Biol. Lett. 2014, 19, 438–460. [Google Scholar] [CrossRef] [PubMed]
  35. Na, H.J.; Akan, I.; Abramowitz, L.K.; Hanover, J.A. Nutrient-Driven O-GlcNAcylation Controls DNA Damage Repair Signaling and Stem/Progenitor Cell Homeostasis. Cell Rep. 2020, 31, 107632. [Google Scholar] [CrossRef]
  36. Nie, H.; Ju, H.; Fan, J.; Shi, X.; Cheng, Y.; Cang, X.; Zheng, Z.; Duan, X.; Yi, W. O-GlcNAcylation of PGK1 coordinates glycolysis and TCA cycle to promote tumor growth. Nat. Commun. 2020, 11, 36. [Google Scholar] [CrossRef]
  37. Xu, W.; Zhang, X.; Wu, J.L.; Fu, L.; Liu, K.; Liu, D.; Chen, G.G.; Lai, P.B.; Wong, N.; Yu, J. O-GlcNAc transferase promotes fatty liver-associated liver cancer through inducing palmitic acid and activating endoplasmic reticulum stress. J. Hepatol. 2017, 67, 310–320. [Google Scholar] [CrossRef]
  38. Liu, Y.; Chen, Q.; Zhang, N.; Zhang, K.; Dou, T.; Cao, Y.; Liu, Y.; Li, K.; Hao, X.; Xie, X.; et al. Proteomic profiling and genome-wide mapping of O-GlcNAc chromatin-associated proteins reveal an O-GlcNAc-regulated genotoxic stress response. Nat. Commun. 2020, 11, 5898. [Google Scholar] [CrossRef]
  39. Schwein, P.A.; Woo, C.M. The O-GlcNAc Modification on Kinases. ACS Chem. Biol. 2020, 15, 602–617. [Google Scholar] [CrossRef]
  40. Hart, G.W.; Slawson, C.; Ramirez-Correa, G.; Lagerlof, O. Cross talk between O-GlcNAcylation and phosphorylation: Roles in signaling, transcription, and chronic disease. Annu. Rev. Biochem. 2011, 80, 825–858. [Google Scholar] [CrossRef] [Green Version]
  41. Butkinaree, C.; Park, K.; Hart, G.W. O-linked beta-N-acetylglucosamine (O-GlcNAc): Extensive crosstalk with phosphorylation to regulate signaling and transcription in response to nutrients and stress. Biochim. Biophys. Acta 2010, 1800, 96–106. [Google Scholar] [CrossRef] [PubMed]
  42. Lee, B.E.; Suh, P.G.; Kim, J.I. O-GlcNAcylation in health and neurodegenerative diseases. Exp. Mol. Med. 2021, 53, 1674–1682. [Google Scholar] [CrossRef]
  43. Aulak, K.S.; Barnes, J.W.; Tian, L.; Mellor, N.E.; Haque, M.M.; Willard, B.; Li, L.; Comhair, S.C.; Stuehr, D.J.; Dweik, R.A. Specific O-GlcNAc modification at Ser-615 modulates eNOS function. Redox Biol. 2020, 36, 101625. [Google Scholar] [CrossRef] [PubMed]
  44. Umapathi, P.; Mesubi, O.O.; Banerjee, P.S.; Abrol, N.; Wang, Q.; Luczak, E.D.; Wu, Y.; Granger, J.M.; Wei, A.C.; Gaido, O.E.R.; et al. Excessive O-GlcNAcylation Causes Heart Failure and Sudden Death. Circulation 2021, 143, 1687–1703. [Google Scholar] [CrossRef] [PubMed]
  45. Banerjee, P.S.; Ma, J.; Hart, G.W. Diabetes-associated dysregulation of O-GlcNAcylation in rat cardiac mitochondria. Proc. Natl. Acad. Sci. USA 2015, 112, 6050–6055. [Google Scholar] [CrossRef] [PubMed]
  46. Yang, Y.; Fu, M.; Li, M.D.; Zhang, K.; Zhang, B.; Wang, S.; Liu, Y.; Ni, W.; Ong, Q.; Mi, J.; et al. O-GlcNAc transferase inhibits visceral fat lipolysis and promotes diet-induced obesity. Nat. Commun. 2020, 11, 181. [Google Scholar] [CrossRef]
  47. Petrus, P.; Lecoutre, S.; Dollet, L.; Wiel, C.; Sulen, A.; Gao, H.; Tavira, B.; Laurencikiene, J.; Rooyackers, O.; Checa, A.; et al. Glutamine Links Obesity to Inflammation in Human White Adipose Tissue. Cell Metab. 2020, 31, 375–390.e311. [Google Scholar] [CrossRef]
  48. Liu, Y.; Hu, Y.J.; Fan, W.X.; Quan, X.; Xu, B.; Li, S.Z. O-GlcNAcylation: The Underestimated Emerging Regulators of Skeletal Muscle Physiology. Cells 2022, 11, 1789. [Google Scholar] [CrossRef]
  49. Ciraku, L.; Bacigalupa, Z.A.; Ju, J.; Moeller, R.A.; Le Minh, G.; Lee, R.H.; Smith, M.D.; Ferrer, C.M.; Trefely, S.; Izzo, L.T.; et al. O-GlcNAc transferase regulates glioblastoma acetate metabolism via regulation of CDK5-dependent ACSS2 phosphorylation. Oncogene 2022, 41, 2122–2136. [Google Scholar] [CrossRef]
  50. Lu, S.; Liao, Z.; Lu, X.; Katschinski, D.M.; Mercola, M.; Chen, J.; Brown, J.H.; Molkentin, J.D.; Bossuyt, J.; Bers, D.M. Hyperglycemia Acutely Increases Cytosolic Reactive Oxygen Species via O-linked GlcNAcylation and CaMKII Activation in Mouse Ventricular Myocytes. Circ. Res. 2020, 126, e80–e96. [Google Scholar] [CrossRef]
  51. Pedowitz, N.J.; Batt, A.R.; Darabedian, N.; Pratt, M.R. MYPT1 O-GlcNAc modification regulates sphingosine-1-phosphate mediated contraction. Nat. Chem. Biol. 2021, 17, 169–177. [Google Scholar] [CrossRef]
  52. Akimoto, Y.; Yan, K.; Miura, Y.; Tsumoto, H.; Toda, T.; Fukutomi, T.; Sugahara, D.; Kudo, A.; Arai, T.; Chiba, Y.; et al. O-GlcNAcylation and phosphorylation of β-actin Ser(199) in diabetic nephropathy. Am. J. Physiol. Ren. Physiol. 2019, 317, F1359–F1374. [Google Scholar] [CrossRef] [PubMed]
  53. Leung, M.C.; Hitchen, P.G.; Ward, D.G.; Messer, A.E.; Marston, S.B. Z-band alternatively spliced PDZ motif protein (ZASP) is the major O-linked β-N-acetylglucosamine-substituted protein in human heart myofibrils. J. Biol. Chem. 2013, 288, 4891–4898. [Google Scholar] [CrossRef] [PubMed]
  54. Hédou, J.; Bastide, B.; Page, A.; Michalski, J.C.; Morelle, W. Mapping of O-linked beta-N-acetylglucosamine modification sites in key contractile proteins of rat skeletal muscle. Proteomics 2009, 9, 2139–2148. [Google Scholar] [CrossRef]
  55. Basu, H.; Pekkurnaz, G.; Falk, J.; Wei, W.; Chin, M.; Steen, J.; Schwarz, T.L. FHL2 anchors mitochondria to actin and adapts mitochondrial dynamics to glucose supply. J. Cell Biol. 2021, 220, e201912077. [Google Scholar] [CrossRef]
  56. Ramirez-Correa, G.A.; Jin, W.; Wang, Z.; Zhong, X.; Gao, W.D.; Dias, W.B.; Vecoli, C.; Hart, G.W.; Murphy, A.M. O-linked GlcNAc modification of cardiac myofilament proteins: A novel regulator of myocardial contractile function. Circ. Res. 2008, 103, 1354–1358. [Google Scholar] [CrossRef] [PubMed]
  57. Hu, Y.; Liu, Y.; Yang, Y.; Lv, H.; Lian, S.; Xu, B.; Li, S. OGT upregulates myogenic IL-6 by mediating O-GlcNAcylation of p65 in mouse skeletal muscle under cold exposure. J. Cell. Physiol. 2022, 237, 1341–1352. [Google Scholar] [CrossRef]
  58. Huang, P.; Ho, S.R.; Wang, K.; Roessler, B.C.; Zhang, F.; Hu, Y.; Bowe, D.B.; Kudlow, J.E.; Paterson, A.J. Muscle-specific overexpression of NCOATGK, splice variant of O-GlcNAcase, induces skeletal muscle atrophy. Am. J. Physiol. Cell Physiol. 2011, 300, C456–C465. [Google Scholar] [CrossRef]
  59. Stevens, L.; Bastide, B.; Hedou, J.; Cieniewski-Bernard, C.; Montel, V.; Cochon, L.; Dupont, E.; Mounier, Y. Potential regulation of human muscle plasticity by MLC2 post-translational modifications during bed rest and countermeasures. Arch. Biochem. Biophys. 2013, 540, 125–132. [Google Scholar] [CrossRef]
  60. Myslicki, J.P.; Belke, D.D.; Shearer, J. Role of O-GlcNAcylation in nutritional sensing, insulin resistance and in mediating the benefits of exercise. Appl. Physiol. Nutr. Metab. Physiol. Appl. Nutr. Metab. 2014, 39, 1205–1213. [Google Scholar] [CrossRef]
  61. Lipina, C.; Hundal, H.S. Lipid modulation of skeletal muscle mass and function. J. Cachexia Sarcopenia Muscle 2017, 8, 190–201. [Google Scholar] [CrossRef] [PubMed]
  62. Shamim, B.; Hawley, J.A.; Camera, D.M. Protein Availability and Satellite Cell Dynamics in Skeletal Muscle. Sports Med. 2018, 48, 1329–1343. [Google Scholar] [CrossRef] [PubMed]
  63. McGee, S.L.; Hargreaves, M. AMPK-mediated regulation of transcription in skeletal muscle. Clin. Sci. 2010, 118, 507–518. [Google Scholar] [CrossRef] [PubMed]
  64. Egan, B.; Zierath, J.R. Exercise metabolism and the molecular regulation of skeletal muscle adaptation. Cell Metab. 2013, 17, 162–184. [Google Scholar] [CrossRef] [PubMed]
  65. Mengeste, A.M.; Rustan, A.C.; Lund, J. Skeletal muscle energy metabolism in obesity. Obesity 2021, 29, 1582–1595. [Google Scholar] [CrossRef] [PubMed]
  66. Henselmans, M.; Bjørnsen, T.; Hedderman, R.; Vårvik, F.T. The Effect of Carbohydrate Intake on Strength and Resistance Training Performance: A Systematic Review. Nutrients 2022, 14, 856. [Google Scholar] [CrossRef] [PubMed]
  67. Yang, J. Enhanced skeletal muscle for effective glucose homeostasis. Prog. Mol. Biol. Transl. Sci. 2014, 121, 133–163. [Google Scholar] [CrossRef]
  68. Giudice, J.; Taylor, J.M. Muscle as a paracrine and endocrine organ. Curr. Opin. Pharmacol. 2017, 34, 49–55. [Google Scholar] [CrossRef]
  69. Pedersen, B.K.; Febbraio, M.A. Muscles, exercise and obesity: Skeletal muscle as a secretory organ. Nat. Rev. Endocrinol. 2012, 8, 457–465. [Google Scholar] [CrossRef]
  70. Singh, J.P.; Qian, K.; Lee, J.S.; Zhou, J.; Han, X.; Zhang, B.; Ong, Q.; Ni, W.; Jiang, M.; Ruan, H.B.; et al. O-GlcNAcase targets pyruvate kinase M2 to regulate tumor growth. Oncogene 2020, 39, 560–573. [Google Scholar] [CrossRef]
  71. Yi, W.; Clark, P.M.; Mason, D.E.; Keenan, M.C.; Hill, C.; Goddard, W.A., 3rd; Peters, E.C.; Driggers, E.M.; Hsieh-Wilson, L.C. Phosphofructokinase 1 glycosylation regulates cell growth and metabolism. Science 2012, 337, 975–980. [Google Scholar] [CrossRef] [PubMed]
  72. 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. Skeletal muscle O-GlcNAc transferase is important for muscle energy homeostasis and whole-body insulin sensitivity. Mol. Metab. 2018, 11, 160–177. [Google Scholar] [CrossRef] [PubMed]
  73. Durning, S.P.; Flanagan-Steet, H.; Prasad, N.; Wells, L. O-Linked β-N-acetylglucosamine (O-GlcNAc) Acts as a Glucose Sensor to Epigenetically Regulate the Insulin Gene in Pancreatic Beta Cells. J. Biol. Chem. 2016, 291, 2107–2118. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, Y.; Xu, B.; Hu, Y.; Liu, P.; Lian, S.; Lv, H.; Yang, Y.; Ji, H.; Yang, H.; Liu, J.; et al. O-GlcNAc/Akt pathway regulates glucose metabolism and reduces apoptosis in liver of piglets with acute cold stress. Cryobiology 2021, 100, 125–132. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, Z.; Pandey, A.; Hart, G.W. Dynamic interplay between O-linked N-acetylglucosaminylation and glycogen synthase kinase-3-dependent phosphorylation. Mol. Cell. Proteom. MCP 2007, 6, 1365–1379. [Google Scholar] [CrossRef]
  76. Hanover, J.A.; Krause, M.W.; Love, D.C. The hexosamine signaling pathway: O-GlcNAc cycling in feast or famine. Biochim. Biophys. Acta 2010, 1800, 80–95. [Google Scholar] [CrossRef]
  77. Filhoulaud, G.; Guillemain, G.; Scharfmann, R. The hexosamine biosynthesis pathway is essential for pancreatic beta cell development. J. Biol. Chem. 2009, 284, 24583–24594. [Google Scholar] [CrossRef]
  78. Richter, E.A.; Hargreaves, M. Exercise, GLUT4, and skeletal muscle glucose uptake. Physiol. Rev. 2013, 93, 993–1017. [Google Scholar] [CrossRef]
  79. Love, D.C.; Hanover, J.A. The hexosamine signaling pathway: Deciphering the “O-GlcNAc code”. Sci. STKE Signal Transduct. Knowl. Environ. 2005, 2005, re13. [Google Scholar] [CrossRef]
  80. Ruegenberg, S.; Horn, M.; Pichlo, C.; Allmeroth, K.; Baumann, U.; Denzel, M.S. Loss of GFAT-1 feedback regulation activates the hexosamine pathway that modulates protein homeostasis. Nat. Commun. 2020, 11, 687. [Google Scholar] [CrossRef] [Green Version]
  81. Campbell, S.; Mesaros, C.; Izzo, L.; Affronti, H.; Noji, M.; Schaffer, B.E.; Tsang, T.; Sun, K.; Trefely, S.; Kruijning, S.; et al. Glutamine deprivation triggers NAGK-dependent hexosamine salvage. eLife 2021, 10, e62644. [Google Scholar] [CrossRef]
  82. Horn, M.; Denzel, S.I.; Srinivasan, B.; Allmeroth, K.; Schiffer, I.; Karthikaisamy, V.; Miethe, S.; Breuer, P.; Antebi, A.; Denzel, M.S. Hexosamine Pathway Activation Improves Protein Homeostasis through the Integrated Stress Response. iScience 2020, 23, 100887. [Google Scholar] [CrossRef] [PubMed]
  83. Lam, C.; Low, J.Y.; Tran, P.T.; Wang, H. The hexosamine biosynthetic pathway and cancer: Current knowledge and future therapeutic strategies. Cancer Lett. 2021, 503, 11–18. [Google Scholar] [CrossRef]
  84. Zachara, N.E.; Hart, G.W. O-GlcNAc a sensor of cellular state: The role of nucleocytoplasmic glycosylation in modulating cellular function in response to nutrition and stress. Biochim. Biophys. Acta 2004, 1673, 13–28. [Google Scholar] [CrossRef]
  85. Zhao, L.; Shah, J.A.; Cai, Y.; Jin, J. ‘O-GlcNAc Code’ Mediated Biological Functions of Downstream Proteins. Molecules 2018, 23, 1967. [Google Scholar] [CrossRef]
  86. Chatham, J.C.; Nöt, L.G.; Fülöp, N.; Marchase, R.B. Hexosamine biosynthesis and protein O-glycosylation: The first line of defense against stress, ischemia, and trauma. Shock 2008, 29, 431–440. [Google Scholar] [CrossRef]
  87. Groves, J.A.; Lee, A.; Yildirir, G.; Zachara, N.E. Dynamic O-GlcNAcylation and its roles in the cellular stress response and homeostasis. Cell Stress Chaperones 2013, 18, 535–558. [Google Scholar] [CrossRef]
  88. Lambert, M.; Bastide, B.; Cieniewski-Bernard, C. Involvement of O-GlcNAcylation in the Skeletal Muscle Physiology and Physiopathology: Focus on Muscle Metabolism. Front. Endocrinol. 2018, 9, 578. [Google Scholar] [CrossRef]
  89. Wang, X.; Feng, Z.; Wang, X.; Yang, L.; Han, S.; Cao, K.; Xu, J.; Zhao, L.; Zhang, Y.; Liu, J. O-GlcNAcase deficiency suppresses skeletal myogenesis and insulin sensitivity in mice through the modulation of mitochondrial homeostasis. Diabetologia 2016, 59, 1287–1296. [Google Scholar] [CrossRef]
  90. Guo, X.; Sun, W.; Luo, G.; Wu, L.; Xu, G.; Hou, D.; Hou, Y.; Guo, X.; Mu, X.; Qin, L.; et al. Panax notoginseng saponins alleviate skeletal muscle insulin resistance by regulating the IRS1-PI3K-AKT signaling pathway and GLUT4 expression. FEBS Open Bio 2019, 9, 1008–1019. [Google Scholar] [CrossRef] [Green Version]
  91. Hawkins, M.; Angelov, I.; Liu, R.; Barzilai, N.; Rossetti, L. The tissue concentration of UDP-N-acetylglucosamine modulates the stimulatory effect of insulin on skeletal muscle glucose uptake. J. Biol. Chem. 1997, 272, 4889–4895. [Google Scholar] [CrossRef] [PubMed]
  92. Ferrer, C.M.; Lynch, T.P.; Sodi, V.L.; Falcone, J.N.; Schwab, L.P.; Peacock, D.L.; Vocadlo, D.J.; Seagroves, T.N.; Reginato, M.J. O-GlcNAcylation regulates cancer metabolism and survival stress signaling via regulation of the HIF-1 pathway. Mol. Cell 2014, 54, 820–831. [Google Scholar] [CrossRef] [PubMed]
  93. Baldini, S.F.; Steenackers, A.; Stichelen, S.O.-V.; Mir, A.M.; Mortuaire, M.; Lefebvre, T.; Guinez, C. Glucokinase expression is regulated by glucose through O-GlcNAc glycosylation. Biochem. Biophys. Res. Commun. 2016, 478, 942–948. [Google Scholar] [CrossRef] [PubMed]
  94. Rao, X.; Duan, X.; Mao, W.; Li, X.; Li, Z.; Li, Q.; Zheng, Z.; Xu, H.; Chen, M.; Wang, P.G.; et al. O-GlcNAcylation of G6PD promotes the pentose phosphate pathway and tumor growth. Nat. Commun. 2015, 6, 8468. [Google Scholar] [CrossRef]
  95. Lei, Y.; Chen, T.; Li, Y.; Shang, M.; Zhang, Y.; Jin, Y.; Yu, Q.; Guo, F.; Wang, T. O-GlcNAcylation of PFKFB3 is required for tumor cell proliferation under hypoxia. Oncogenesis 2020, 9, 21. [Google Scholar] [CrossRef]
  96. Lim, K.; Yoon, B.H.; Ha, C.H. O-Linked N-acetylglucosaminylation of Sp1 interferes with Sp1 activation of glycolytic genes. Biochem. Biophys. Res. Commun. 2015, 468, 349–353. [Google Scholar] [CrossRef]
  97. Bacigalupa, Z.A.; Bhadiadra, C.H.; Reginato, M.J. O-GlcNAcylation: Key regulator of glycolytic pathways. J. Bioenerg. Biomembr. 2018, 50, 189–198. [Google Scholar] [CrossRef]
  98. Nachit, M.; Leclercq, I.A. Emerging awareness on the importance of skeletal muscle in liver diseases: Time to dig deeper into mechanisms! Clin. Sci. 2019, 133, 465–481. [Google Scholar] [CrossRef]
  99. Kazemi, Z.; Chang, H.; Haserodt, S.; McKen, C.; Zachara, N.E. O-linked beta-N-acetylglucosamine (O-GlcNAc) regulates stress-induced heat shock protein expression in a GSK-3beta-dependent manner. J. Biol. Chem. 2010, 285, 39096–39107. [Google Scholar] [CrossRef] [PubMed]
  100. Parker, G.J.; Lund, K.C.; Taylor, R.P.; McClain, D.A. Insulin resistance of glycogen synthase mediated by o-linked N-acetylglucosamine. J. Biol. Chem. 2003, 278, 10022–10027. [Google Scholar] [CrossRef] [Green Version]
  101. Zumbaugh, M.D.; Yen, C.N.; Bodmer, J.S.; Shi, H.; Gerrard, D.E. Skeletal Muscle O-GlcNAc Transferase Action on Global Metabolism is Partially Mediated Through Interleukin-15. Front. Physiol. 2021, 12, 682052. [Google Scholar] [CrossRef] [PubMed]
  102. Laczy, B.; Fülöp, N.; Onay-Besikci, A.; Des Rosiers, C.; Chatham, J.C. Acute regulation of cardiac metabolism by the hexosamine biosynthesis pathway and protein O-GlcNAcylation. PLoS ONE 2011, 6, e18417. [Google Scholar] [CrossRef] [PubMed]
  103. Guinez, C.; Filhoulaud, G.; Rayah-Benhamed, F.; Marmier, S.; Dubuquoy, C.; Dentin, R.; Moldes, M.; Burnol, A.F.; Yang, X.; Lefebvre, T.; et al. O-GlcNAcylation increases ChREBP protein content and transcriptional activity in the liver. Diabetes 2011, 60, 1399–1413. [Google Scholar] [CrossRef] [PubMed]
  104. Sodi, V.L.; Bacigalupa, Z.A.; Ferrer, C.M.; Lee, J.V.; Gocal, W.A.; Mukhopadhyay, D.; Wellen, K.E.; Ivan, M.; Reginato, M.J. Nutrient sensor O-GlcNAc transferase controls cancer lipid metabolism via SREBP-1 regulation. Oncogene 2018, 37, 924–934. [Google Scholar] [CrossRef]
  105. Benhamed, F.; Filhoulaud, G.; Caron, S.; Lefebvre, P.; Staels, B.; Postic, C. O-GlcNAcylation Links ChREBP and FXR to Glucose-Sensing. Front. Endocrinol. 2014, 5, 230. [Google Scholar] [CrossRef]
  106. Anthonisen, E.H.; Berven, L.; Holm, S.; Nygård, M.; Nebb, H.I.; Grønning-Wang, L.M. Nuclear receptor liver X receptor is O-GlcNAc-modified in response to glucose. J. Biol. Chem. 2010, 285, 1607–1615. [Google Scholar] [CrossRef]
  107. Wong, Y.K.; Wang, J.; Lim, T.K.; Lin, Q.; Yap, C.T.; Shen, H.M. O-GlcNAcylation promotes fatty acid synthase activity under nutritional stress as a pro-survival mechanism in cancer cells. Proteomics 2022, 22, e2100175. [Google Scholar] [CrossRef]
  108. Deshmukh, A.S. Insulin-stimulated glucose uptake in healthy and insulin-resistant skeletal muscle. Horm. Mol. Biol. Clin. Investig. 2016, 26, 13–24. [Google Scholar] [CrossRef] [PubMed]
  109. McLarty, J.L.; Marsh, S.A.; Chatham, J.C. Post-translational protein modification by O-linked N-acetyl-glucosamine: Its role in mediating the adverse effects of diabetes on the heart. Life Sci 2013, 92, 621–627. [Google Scholar] [CrossRef]
  110. Wells, L.; Vosseller, K.; Hart, G.W. A role for N-acetylglucosamine as a nutrient sensor and mediator of insulin resistance. Cell. Mol. Life Sci. CMLS 2003, 60, 222–228. [Google Scholar] [CrossRef]
  111. Virkamäki, A.; Daniels, M.C.; Hämäläinen, S.; Utriainen, T.; McClain, D.; Yki-Järvinen, H. Activation of the hexosamine pathway by glucosamine in vivo induces insulin resistance in multiple insulin sensitive tissues. Endocrinology 1997, 138, 2501–2507. [Google Scholar] [CrossRef]
  112. Arias, E.B.; Kim, J.; Cartee, G.D. Prolonged incubation in PUGNAc results in increased protein O-Linked glycosylation and insulin resistance in rat skeletal muscle. Diabetes 2004, 53, 921–930. [Google Scholar] [CrossRef] [PubMed]
  113. Walgren, J.L.; Vincent, T.S.; Schey, K.L.; Buse, M.G. High glucose and insulin promote O-GlcNAc modification of proteins, including alpha-tubulin. Am. J. Physiol. Endocrinol. Metab. 2003, 284, E424–E434. [Google Scholar] [CrossRef] [PubMed]
  114. Yki-Järvinen, H.; Virkamäki, A.; Daniels, M.C.; McClain, D.; Gottschalk, W.K. Insulin and glucosamine infusions increase O-linked N-acetyl-glucosamine in skeletal muscle proteins in vivo. Metab. Clin. Exp. 1998, 47, 449–455. [Google Scholar] [CrossRef]
  115. Ansari, S.A.; Emerald, B.S. The Role of Insulin Resistance and Protein O-GlcNAcylation in Neurodegeneration. Front. Neurosci. 2019, 13, 473. [Google Scholar] [CrossRef] [PubMed]
  116. Copeland, R.J.; Bullen, J.W.; Hart, G.W. Cross-talk between GlcNAcylation and phosphorylation: Roles in insulin resistance and glucose toxicity. Am. J. Physiol. Endocrinol. Metab. 2008, 295, E17–E28. [Google Scholar] [CrossRef]
  117. Rojas, F.A.; Hirata, A.E.; Saad, M.J. Regulation of insulin receptor substrate-2 tyrosine phosphorylation in animal models of insulin resistance. Endocrine 2003, 21, 115–122. [Google Scholar] [CrossRef]
  118. Copps, K.D.; White, M.F. Regulation of insulin sensitivity by serine/threonine phosphorylation of insulin receptor substrate proteins IRS1 and IRS2. Diabetologia 2012, 55, 2565–2582. [Google Scholar] [CrossRef]
  119. Yudushkin, I. Getting the Akt Together: Guiding Intracellular Akt Activity by PI3K. Biomolecules 2019, 9, 67. [Google Scholar] [CrossRef]
  120. Ebner, M.; Lučić, I.; Leonard, T.A.; Yudushkin, I. PI(3,4,5)P(3) Engagement Restricts Akt Activity to Cellular Membranes. Mol. Cell 2017, 65, 416–431.e416. [Google Scholar] [CrossRef] [Green Version]
  121. Wick, K.L.; Liu, F. A new molecular target of insulin action: Regulating the pivotal PDK1. Curr. Drug Targets Immune Endocr. Metab. Disord. 2001, 1, 209–221. [Google Scholar] [CrossRef] [PubMed]
  122. Yudushkin, I. Control of Akt activity and substrate phosphorylation in cells. IUBMB Life 2020, 72, 1115–1125. [Google Scholar] [CrossRef] [PubMed]
  123. Goldbraikh, D.; Neufeld, D.; Eid-Mutlak, Y.; Lasry, I.; Gilda, J.E.; Parnis, A.; Cohen, S. USP1 deubiquitinates Akt to inhibit PI3K-Akt-FoxO signaling in muscle during prolonged starvation. EMBO Rep. 2020, 21, e48791. [Google Scholar] [CrossRef]
  124. Xie, Y.; Shi, X.; Sheng, K.; Han, G.; Li, W.; Zhao, Q.; Jiang, B.; Feng, J.; Li, J.; Gu, Y. PI3K/Akt signaling transduction pathway, erythropoiesis and glycolysis in hypoxia (Review). Mol. Med. Rep. 2019, 19, 783–791. [Google Scholar] [CrossRef] [PubMed]
  125. Camaforte, N.A.P.; Saldanha, L.L.; Vareda, P.M.P.; Rezende-Neto, J.M.; Senger, M.R.; Delgado, A.Q.; Morgan, H.J.N.; Violato, N.M.; Pieroni, L.G.; Dokkedal, A.L.; et al. Hypoglycaemic activity of Bauhinia holophylla through GSK3-β inhibition and glycogenesis activation. Pharm. Biol. 2019, 57, 269–279. [Google Scholar] [CrossRef] [PubMed]
  126. Yang, X.; Ongusaha, P.P.; Miles, P.D.; Havstad, J.C.; Zhang, F.; So, W.V.; Kudlow, J.E.; Michell, R.H.; Olefsky, J.M.; Field, S.J.; et al. Phosphoinositide signalling links O-GlcNAc transferase to insulin resistance. Nature 2008, 451, 964–969. [Google Scholar] [CrossRef]
  127. Whelan, S.A.; Lane, M.D.; Hart, G.W. Regulation of the O-linked beta-N-acetylglucosamine transferase by insulin signaling. J. Biol. Chem. 2008, 283, 21411–21417. [Google Scholar] [CrossRef] [PubMed]
  128. Whelan, S.A.; Dias, W.B.; Thiruneelakantapillai, L.; Lane, M.D.; Hart, G.W. Regulation of insulin receptor substrate 1 (IRS-1)/AKT kinase-mediated insulin signaling by O-Linked beta-N-acetylglucosamine in 3T3-L1 adipocytes. J. Biol. Chem. 2010, 285, 5204–5211. [Google Scholar] [CrossRef]
  129. Issad, T.; Masson, E.; Pagesy, P. O-GlcNAc modification, insulin signaling and diabetic complications. Diabetes Metab. 2010, 36, 423–435. [Google Scholar] [CrossRef]
  130. Vosseller, K.; Wells, L.; Lane, M.D.; Hart, G.W. Elevated nucleocytoplasmic glycosylation by O-GlcNAc results in insulin resistance associated with defects in Akt activation in 3T3-L1 adipocytes. Proc. Natl. Acad. Sci. USA 2002, 99, 5313–5318. [Google Scholar] [CrossRef] [Green Version]
  131. Zhang, Z.Y.; Dodd, G.T.; Tiganis, T. Protein Tyrosine Phosphatases in Hypothalamic Insulin and Leptin Signaling. Trends Pharmacol. Sci. 2015, 36, 661–674. [Google Scholar] [CrossRef] [PubMed]
  132. Zhao, Y.; Tang, Z.; Shen, A.; Tao, T.; Wan, C.; Zhu, X.; Huang, J.; Zhang, W.; Xia, N.; Wang, S.; et al. The Role of PTP1B O-GlcNAcylation in Hepatic Insulin Resistance. Int. J. Mol. Sci. 2015, 16, 22856–22869. [Google Scholar] [CrossRef] [PubMed]
  133. Yang, Y.; Li, X.; Luan, H.H.; Zhang, B.; Zhang, K.; Nam, J.H.; Li, Z.; Fu, M.; Munk, A.; Zhang, D.; et al. OGT suppresses S6K1-mediated macrophage inflammation and metabolic disturbance. Proc. Natl. Acad. Sci. USA 2020, 117, 16616–16625. [Google Scholar] [CrossRef]
  134. Gao, Y.; Liu, J.; Bai, Z.; Sink, S.; Zhao, C.; Lorenzo, F.R.; McClain, D.A. Iron down-regulates leptin by suppressing protein O-GlcNAc modification in adipocytes, resulting in decreased levels of O-glycosylated CREB. J. Biol. Chem. 2019, 294, 5487–5495. [Google Scholar] [CrossRef]
  135. Rexach, J.E.; Clark, P.M.; Mason, D.E.; Neve, R.L.; Peters, E.C.; Hsieh-Wilson, L.C. Dynamic O-GlcNAc modification regulates CREB-mediated gene expression and memory formation. Nat. Chem. Biol. 2012, 8, 253–261. [Google Scholar] [CrossRef]
  136. Dentin, R.; Hedrick, S.; Xie, J.; Yates, J., 3rd; Montminy, M. Hepatic glucose sensing via the CREB coactivator CRTC2. Science 2008, 319, 1402–1405. [Google Scholar] [CrossRef]
  137. Romanello, V.; Sandri, M. The connection between the dynamic remodeling of the mitochondrial network and the regulation of muscle mass. Cell. Mol. Life Sci. CMLS 2021, 78, 1305–1328. [Google Scholar] [CrossRef]
  138. Parry, H.A.; Roberts, M.D.; Kavazis, A.N. Human Skeletal Muscle Mitochondrial Adaptations Following Resistance Exercise Training. Int. J. Sports Med. 2020, 41, 349–359. [Google Scholar] [CrossRef]
  139. Banerjee, P.S.; Lagerlöf, O.; Hart, G.W. Roles of O-GlcNAc in chronic diseases of aging. Mol. Asp. Med. 2016, 51, 1–15. [Google Scholar] [CrossRef]
  140. Wang, Z.; Qin, J.; Zhao, J.; Li, J.; Li, D.; Popp, M.; Popp, F.; Alakus, H.; Kong, B.; Dong, Q.; et al. Inflammatory IFIT3 renders chemotherapy resistance by regulating post-translational modification of VDAC2 in pancreatic cancer. Theranostics 2020, 10, 7178–7192. [Google Scholar] [CrossRef]
  141. Hanover, J.A.; Yu, S.; Lubas, W.B.; Shin, S.H.; Ragano-Caracciola, M.; Kochran, J.; Love, D.C. Mitochondrial and nucleocytoplasmic isoforms of O-linked GlcNAc transferase encoded by a single mammalian gene. Arch. Biochem. Biophys. 2003, 409, 287–297. [Google Scholar] [CrossRef]
  142. Hu, Y.; Suarez, J.; Fricovsky, E.; Wang, H.; Scott, B.T.; Trauger, S.A.; Han, W.; Hu, Y.; Oyeleye, M.O.; Dillmann, W.H. Increased enzymatic O-GlcNAcylation of mitochondrial proteins impairs mitochondrial function in cardiac myocytes exposed to high glucose. J. Biol. Chem. 2009, 284, 547–555. [Google Scholar] [CrossRef] [PubMed]
  143. Cao, W.; Cao, J.; Huang, J.; Yao, J.; Yan, G.; Xu, H.; Yang, P. Discovery and confirmation of O-GlcNAcylated proteins in rat liver mitochondria by combination of mass spectrometry and immunological methods. PLoS ONE 2013, 8, e76399. [Google Scholar] [CrossRef]
  144. Zumbaugh, M.D.; Johnson, S.E.; Shi, T.H.; Gerrard, D.E. Molecular and biochemical regulation of skeletal muscle metabolism. J. Anim. Sci. 2022, 100, skac035. [Google Scholar] [CrossRef] [PubMed]
  145. Ma, J.; Banerjee, P.; Whelan, S.A.; Liu, T.; Wei, A.C.; Ramirez-Correa, G.; McComb, M.E.; Costello, C.E.; O’Rourke, B.; Murphy, A.; et al. Comparative Proteomics Reveals Dysregulated Mitochondrial O-GlcNAcylation in Diabetic Hearts. J. Proteome Res. 2016, 15, 2254–2264. [Google Scholar] [CrossRef] [PubMed]
  146. Ma, J.; Liu, T.; Wei, A.C.; Banerjee, P.; O’Rourke, B.; Hart, G.W. O-GlcNAcomic Profiling Identifies Widespread O-Linked β-N-Acetylglucosamine Modification (O-GlcNAcylation) in Oxidative Phosphorylation System Regulating Cardiac Mitochondrial Function. J. Biol. Chem. 2015, 290, 29141–29153. [Google Scholar] [CrossRef]
  147. Wu, J.; Liu, J.; Lapenta, K.; Desrouleaux, R.; Li, M.D.; Yang, X. Regulation of the urea cycle by CPS1 O-GlcNAcylation in response to dietary restriction and aging. J. Mol. Cell Biol. 2022, 14, mjac016. [Google Scholar] [CrossRef]
  148. Zhao, L.; Feng, Z.; Yang, X.; Liu, J. The regulatory roles of O-GlcNAcylation in mitochondrial homeostasis and metabolic syndrome. Free Radic. Res. 2016, 50, 1080–1088. [Google Scholar] [CrossRef]
  149. Ohashi, N.; Morino, K.; Ida, S.; Sekine, O.; Lemecha, M.; Kume, S.; Park, S.Y.; Choi, C.S.; Ugi, S.; Maegawa, H. Pivotal Role of O-GlcNAc Modification in Cold-Induced Thermogenesis by Brown Adipose Tissue Through Mitochondrial Biogenesis. Diabetes 2017, 66, 2351–2362. [Google Scholar] [CrossRef]
  150. Pekkurnaz, G.; Trinidad, J.C.; Wang, X.; Kong, D.; Schwarz, T.L. Glucose regulates mitochondrial motility via Milton modification by O-GlcNAc transferase. Cell 2014, 158, 54–68. [Google Scholar] [CrossRef] [Green Version]
  151. Gawlowski, T.; Suarez, J.; Scott, B.; Torres-Gonzalez, M.; Wang, H.; Schwappacher, R.; Han, X.; Yates, J.R., 3rd; Hoshijima, M.; Dillmann, W. Modulation of dynamin-related protein 1 (DRP1) function by increased O-linked-β-N-acetylglucosamine modification (O-GlcNAc) in cardiac myocytes. J. Biol. Chem. 2012, 287, 30024–30034. [Google Scholar] [CrossRef] [PubMed]
  152. Akinbiyi, E.O.; Abramowitz, L.K.; Bauer, B.L.; Stoll, M.S.K.; Hoppel, C.L.; Hsiao, C.P.; Hanover, J.A.; Mears, J.A. Blocked O-GlcNAc cycling alters mitochondrial morphology, function, and mass. Sci. Rep. 2021, 11, 22106. [Google Scholar] [CrossRef] [PubMed]
  153. Jin, J.Y.; Wei, X.X.; Zhi, X.L.; Wang, X.H.; Meng, D. Drp1-dependent mitochondrial fission in cardiovascular disease. Acta Pharmacol. Sin. 2021, 42, 655–664. [Google Scholar] [CrossRef] [PubMed]
  154. Champattanachai, V.; Marchase, R.B.; Chatham, J.C. Glucosamine protects neonatal cardiomyocytes from ischemia-reperfusion injury via increased protein O-GlcNAc and increased mitochondrial Bcl-2. Am. J. Physiol. Cell Physiol. 2008, 294, C1509–C1520. [Google Scholar] [CrossRef] [PubMed]
  155. Marsh, S.A.; Powell, P.C.; Dell’italia, L.J.; Chatham, J.C. Cardiac O-GlcNAcylation blunts autophagic signaling in the diabetic heart. Life Sci. 2013, 92, 648–656. [Google Scholar] [CrossRef]
  156. Cividini, F.; Scott, B.T.; Dai, A.; Han, W.; Suarez, J.; Diaz-Juarez, J.; Diemer, T.; Casteel, D.E.; Dillmann, W.H. O-GlcNAcylation of 8-Oxoguanine DNA Glycosylase (Ogg1) Impairs Oxidative Mitochondrial DNA Lesion Repair in Diabetic Hearts. J. Biol. Chem. 2016, 291, 26515–26528. [Google Scholar] [CrossRef]
  157. Jones, S.P.; Zachara, N.E.; Ngoh, G.A.; Hill, B.G.; Teshima, Y.; Bhatnagar, A.; Hart, G.W.; Marbán, E. Cardioprotection by N-acetylglucosamine linkage to cellular proteins. Circulation 2008, 117, 1172–1182. [Google Scholar] [CrossRef]
  158. Dinić, S.; Arambašić, J.; Mihailović, M.; Uskoković, A.; Grdović, N.; Marković, J.; Karadžić, B.; Poznanović, G.; Vidaković, M. Decreased O-GlcNAcylation of the key proteins in kinase and redox signalling pathways is a novel mechanism of the beneficial effect of α-lipoic acid in diabetic liver. Br. J. Nutr. 2013, 110, 401–412. [Google Scholar] [CrossRef]
  159. Bottinelli, R.; Reggiani, C. Human skeletal muscle fibres: Molecular and functional diversity. Prog. Biophys. Mol. Biol. 2000, 73, 195–262. [Google Scholar] [CrossRef]
  160. Bottinelli, R. Functional heterogeneity of mammalian single muscle fibres: Do myosin isoforms tell the whole story? Pflug. Arch. Eur. J. Physiol. 2001, 443, 6–17. [Google Scholar] [CrossRef]
  161. Schiaffino, S.; Reggiani, C. Fiber types in mammalian skeletal muscles. Physiol. Rev. 2011, 91, 1447–1531. [Google Scholar] [CrossRef]
  162. Schiaffino, S. Muscle fiber type diversity revealed by anti-myosin heavy chain antibodies. FEBS J. 2018, 285, 3688–3694. [Google Scholar] [CrossRef] [PubMed]
  163. Westerblad, H.; Bruton, J.D.; Katz, A. Skeletal muscle: Energy metabolism, fiber types, fatigue and adaptability. Exp. Cell Res. 2010, 316, 3093–3099. [Google Scholar] [CrossRef] [PubMed]
  164. Frontera, W.R.; Ochala, J. Skeletal muscle: A brief review of structure and function. Calcif. Tissue Int. 2015, 96, 183–195. [Google Scholar] [CrossRef]
  165. Cieniewski-Bernard, C.; Lambert, M.; Dupont, E.; Montel, V.; Stevens, L.; Bastide, B. O-GlcNAcylation, contractile protein modifications and calcium affinity in skeletal muscle. Front. Physiol. 2014, 5, 421. [Google Scholar] [CrossRef] [PubMed]
  166. Lambert, M.; Richard, E.; Duban-Deweer, S.; Krzewinski, F.; Deracinois, B.; Dupont, E.; Bastide, B.; Cieniewski-Bernard, C. O-GlcNAcylation is a key modulator of skeletal muscle sarcomeric morphometry associated to modulation of protein-protein interactions. Biochim. Biophys. Acta 2016, 1860, 2017–2030. [Google Scholar] [CrossRef]
  167. Cieniewski-Bernard, C.; Montel, V.; Stevens, L.; Bastide, B. O-GlcNAcylation, an original modulator of contractile activity in striated muscle. J. Muscle Res. Cell Motil. 2009, 30, 281–287. [Google Scholar] [CrossRef]
  168. Hortemo, K.H.; Lunde, P.K.; Anonsen, J.H.; Kvaløy, H.; Munkvik, M.; Rehn, T.A.; Sjaastad, I.; Lunde, I.G.; Aronsen, J.M.; Sejersted, O.M. Exercise training increases protein O-GlcNAcylation in rat skeletal muscle. Physiol. Rep. 2016, 4, e12896. [Google Scholar] [CrossRef] [PubMed]
  169. Cieniewski-Bernard, C.; Mounier, Y.; Michalski, J.C.; Bastide, B. O-GlcNAc level variations are associated with the development of skeletal muscle atrophy. J. Appl. Physiol. 2006, 100, 1499–1505. [Google Scholar] [CrossRef]
  170. Cieniewski-Bernard, C.; Dupont, E.; Richard, E.; Bastide, B. Phospho-GlcNAc modulation of slow MLC2 during soleus atrophy through a multienzymatic and sarcomeric complex. Pflug. Arch. 2014, 466, 2139–2151. [Google Scholar] [CrossRef]
  171. Fülöp, N.; Feng, W.; Xing, D.; He, K.; Nőt, L.G.; Brocks, C.A.; Marchase, R.B.; Miller, A.P.; Chatham, J.C. Aging leads to increased levels of protein O-linked N-acetylglucosamine in heart, aorta, brain and skeletal muscle in Brown-Norway rats. Biogerontology 2008, 9, 139. [Google Scholar] [CrossRef]
  172. Toivonen, M.H.; Pöllänen, E.; Ahtiainen, M.; Suominen, H.; Taaffe, D.R.; Cheng, S.; Takala, T.; Kujala, U.M.; Tammi, M.I.; Sipilä, S.; et al. OGT and OGA expression in postmenopausal skeletal muscle associates with hormone replacement therapy and muscle cross-sectional area. Exp. Gerontol. 2013, 48, 1501–1504. [Google Scholar] [CrossRef] [PubMed]
  173. Hortemo, K.H.; Aronsen, J.M.; Lunde, I.G.; Sjaastad, I.; Lunde, P.K.; Sejersted, O.M. Exhausting treadmill running causes dephosphorylation of sMLC2 and reduced level of myofilament MLCK2 in slow twitch rat soleus muscle. Physiol. Rep. 2015, 3, e12285. [Google Scholar] [CrossRef] [PubMed]
  174. Peternelj, T.T.; Marsh, S.A.; Strobel, N.A.; Matsumoto, A.; Briskey, D.; Dalbo, V.J.; Tucker, P.S.; Coombes, J.S. Glutathione depletion and acute exercise increase O-GlcNAc protein modification in rat skeletal muscle. Mol. Cell. Biochem. 2015, 400, 265–275. [Google Scholar] [CrossRef] [PubMed]
  175. Nelson, B.A.; Robinson, K.A.; Koning, J.S.; Buse, M.G. Effects of exercise and feeding on the hexosamine biosynthetic pathway in rat skeletal muscle. Am. J. Physiol. 1997, 272, E848–E855. [Google Scholar] [CrossRef]
  176. Murata, K.; Morino, K.; Ida, S.; Ohashi, N.; Lemecha, M.; Park, S.Y.; Ishikado, A.; Kume, S.; Choi, C.S.; Sekine, O.; et al. Lack of O-GlcNAcylation enhances exercise-dependent glucose utilization potentially through AMP-activated protein kinase activation in skeletal muscle. Biochem. Biophys. Res. Commun. 2018, 495, 2098–2104. [Google Scholar] [CrossRef] [PubMed]
  177. Gransee, H.M.; Mantilla, C.B.; Sieck, G.C. Respiratory muscle plasticity. Compr. Physiol. 2012, 2, 1441–1462. [Google Scholar] [CrossRef] [PubMed]
  178. Konopka, A.R.; Harber, M.P. Skeletal muscle hypertrophy after aerobic exercise training. Exerc. Sport Sci. Rev. 2014, 42, 53–61. [Google Scholar] [CrossRef]
  179. Qaisar, R.; Bhaskaran, S.; Van Remmen, H. Muscle fiber type diversification during exercise and regeneration. Free Radic. Biol. Med. 2016, 98, 56–67. [Google Scholar] [CrossRef]
  180. Blaauw, B.; Schiaffino, S.; Reggiani, C. Mechanisms modulating skeletal muscle phenotype. Compr. Physiol. 2013, 3, 1645–1687. [Google Scholar] [CrossRef]
  181. Teixeira, E.; Duarte, J.A. Skeletal Muscle Loading Changes its Regenerative Capacity. Sports Med. 2016, 46, 783–792. [Google Scholar] [CrossRef]
  182. Davies, K.J.; Quintanilha, A.T.; Brooks, G.A.; Packer, L. Free radicals and tissue damage produced by exercise. Biochem. Biophys. Res. Commun. 1982, 107, 1198–1205. [Google Scholar] [CrossRef]
  183. Pette, D.; Staron, R.S. Myosin isoforms, muscle fiber types, and transitions. Microsc. Res. Tech. 2000, 50, 500–509. [Google Scholar] [CrossRef]
  184. Kim, Y.H.; Song, M.; Oh, Y.S.; Heo, K.; Choi, J.W.; Park, J.M.; Kim, S.H.; Lim, S.; Kwon, H.M.; Ryu, S.H.; et al. Inhibition of phospholipase C-beta1-mediated signaling by O-GlcNAc modification. J. Cell. Physiol. 2006, 207, 689–696. [Google Scholar] [CrossRef] [PubMed]
  185. Slawson, C.; Housley, M.P.; Hart, G.W. O-GlcNAc cycling: How a single sugar post-translational modification is changing the way we think about signaling networks. J. Cell. Biochem. 2006, 97, 71–83. [Google Scholar] [CrossRef]
  186. Peternelj, T.T.; Marsh, S.A.; Morais, C.; Small, D.M.; Dalbo, V.J.; Tucker, P.S.; Coombes, J.S. O-GlcNAc protein modification in C2C12 myoblasts exposed to oxidative stress indicates parallels with endogenous antioxidant defense. Biochem. Cell Biol. Biochim. Biol. Cell. 2015, 93, 63–73. [Google Scholar] [CrossRef] [PubMed]
  187. Lima, V.V.; Spitler, K.; Choi, H.; Webb, R.C.; Tostes, R.C. O-GlcNAcylation and oxidation of proteins: Is signalling in the cardiovascular system becoming sweeter? Clin. Sci. 2012, 123, 473–486. [Google Scholar] [CrossRef]
  188. Chen, P.H.; Chi, J.T.; Boyce, M. Functional crosstalk among oxidative stress and O-GlcNAc signaling pathways. Glycobiology 2018, 28, 556–564. [Google Scholar] [CrossRef] [PubMed]
  189. Booth, F.W.; Laye, M.J.; Spangenburg, E.E. Gold standards for scientists who are conducting animal-based exercise studies. J. Appl. Physiol. 2010, 108, 219–221. [Google Scholar] [CrossRef]
  190. Sun, C.; Shang, J.; Yao, Y.; Yin, X.; Liu, M.; Liu, H.; Zhou, Y. O-GlcNAcylation: A bridge between glucose and cell differentiation. J. Cell. Mol. Med. 2016, 20, 769–781. [Google Scholar] [CrossRef] [Green Version]
  191. Sun, C.; Lan, W.; Li, B.; Zuo, R.; Xing, H.; Liu, M.; Li, J.; Yao, Y.; Wu, J.; Tang, Y.; et al. Glucose regulates tissue-specific chondro-osteogenic differentiation of human cartilage endplate stem cells via O-GlcNAcylation of Sox9 and Runx2. Stem Cell Res. Ther. 2019, 10, 357. [Google Scholar] [CrossRef]
  192. Gu, H.; Song, M.; Boonanantanasarn, K.; Baek, K.; Woo, K.M.; Ryoo, H.M.; Baek, J.H. Conditions Inducing Excessive O-GlcNAcylation Inhibit BMP2-Induced Osteogenic Differentiation of C2C12 Cells. Int. J. Mol. Sci. 2018, 19, 202. [Google Scholar] [CrossRef]
  193. Ogawa, M.; Mizofuchi, H.; Kobayashi, Y.; Tsuzuki, G.; Yamamoto, M.; Wada, S.; Kamemura, K. Terminal differentiation program of skeletal myogenesis is negatively regulated by O-GlcNAc glycosylation. Biochim. Biophys. Acta 2012, 1820, 24–32. [Google Scholar] [CrossRef] [PubMed]
  194. Kim, H.B.; Seo, H.G.; Son, S.; Choi, H.; Kim, B.G.; Kweon, T.H.; Kim, S.; Pai, J.; Shin, I.; Yang, W.H.; et al. O-GlcNAcylation of Mef2c regulates myoblast differentiation. Biochem. Biophys. Res. Commun. 2020, 529, 692–698. [Google Scholar] [CrossRef]
  195. Ogawa, M.; Sakakibara, Y.; Kamemura, K. Requirement of decreased O-GlcNAc glycosylation of Mef2D for its recruitment to the myogenin promoter. Biochem. Biophys. Res. Commun. 2013, 433, 558–562. [Google Scholar] [CrossRef]
  196. Anderson, J.E. Key concepts in muscle regeneration: Muscle “cellular ecology” integrates a gestalt of cellular cross-talk, motility, and activity to remodel structure and restore function. Eur. J. Appl. Physiol. 2022, 122, 273–300. [Google Scholar] [CrossRef] [PubMed]
  197. Byon, C.H.; Kim, S.W. Regulatory Effects of O-GlcNAcylation in Vascular Smooth Muscle Cells on Diabetic Vasculopathy. J. Lipid Atheroscler. 2020, 9, 243–254. [Google Scholar] [CrossRef] [PubMed]
  198. Dang, K.; Jiang, S.; Gao, Y.; Qian, A. The role of protein glycosylation in muscle diseases. Mol. Biol. Rep. 2022, 49, 8037–8049. [Google Scholar] [CrossRef]
  199. Nakamura, S.; Nakano, S.; Nishii, M.; Kaneko, S.; Kusaka, H. Localization of O-GlcNAc-modified proteins in neuromuscular diseases. Med. Mol. Morphol. 2012, 45, 86–90. [Google Scholar] [CrossRef]
  200. Shan, X.; Vocadlo, D.J.; Krieger, C. Reduced protein O-glycosylation in the nervous system of the mutant SOD1 transgenic mouse model of amyotrophic lateral sclerosis. Neurosci. Lett. 2012, 516, 296–301. [Google Scholar] [CrossRef]
  201. Lüdemann, N.; Clement, A.; Hans, V.H.; Leschik, J.; Behl, C.; Brandt, R. O-glycosylation of the tail domain of neurofilament protein M in human neurons and in spinal cord tissue of a rat model of amyotrophic lateral sclerosis (ALS). J. Biol. Chem. 2005, 280, 31648–31658. [Google Scholar] [CrossRef] [PubMed]
  202. Chen, P.H.; Hu, J.; Wu, J.; Huynh, D.T.; Smith, T.J.; Pan, S.; Bisnett, B.J.; Smith, A.B.; Lu, A.; Condon, B.M.; et al. Gigaxonin glycosylation regulates intermediate filament turnover and may impact giant axonal neuropathy etiology or treatment. JCI Insight 2020, 5, e127751. [Google Scholar] [CrossRef] [PubMed]
  203. Malicdan, M.C.; Noguchi, S.; Nonaka, I.; Hayashi, Y.K.; Nishino, I. A Gne knockout mouse expressing human GNE D176V mutation develops features similar to distal myopathy with rimmed vacuoles or hereditary inclusion body myopathy. Hum. Mol. Genet. 2007, 16, 2669–2682. [Google Scholar] [CrossRef]
  204. Bennmann, D.; Weidemann, W.; Thate, A.; Kreuzmann, D.; Horstkorte, R. Aberrant O-GlcNAcylation disrupts GNE enzyme activity in GNE myopathy. FEBS J. 2016, 283, 2285–2294. [Google Scholar] [CrossRef]
  205. Massaccesi, L.; Goi, G.; Tringali, C.; Barassi, A.; Venerando, B.; Papini, N. Dexamethasone-Induced Skeletal Muscle Atrophy Increases O-GlcNAcylation in C2C12 Cells. J. Cell. Biochem. 2016, 117, 1833–1842. [Google Scholar] [CrossRef]
  206. Akan, I.; Olivier-Van Stichelen, S.; Bond, M.R.; Hanover, J.A. Nutrient-driven O-GlcNAc in proteostasis and neurodegeneration. J. Neurochem. 2018, 144, 7–34. [Google Scholar] [CrossRef]
  207. Martin-Rincon, M.; Morales-Alamo, D.; Calbet, J.A.L. Exercise-mediated modulation of autophagy in skeletal muscle. Scand. J. Med. Sci. Sports 2018, 28, 772–781. [Google Scholar] [CrossRef]
  208. Rodney, G.G.; Pal, R.; Abo-Zahrah, R. Redox regulation of autophagy in skeletal muscle. Free Radic. Biol. Med. 2016, 98, 103–112. [Google Scholar] [CrossRef] [PubMed]
  209. Jiao, J.; Demontis, F. Skeletal muscle autophagy and its role in sarcopenia and organismal aging. Curr. Opin. Pharmacol. 2017, 34, 1–6. [Google Scholar] [CrossRef]
  210. Xie, Y.; Kang, R.; Sun, X.; Zhong, M.; Huang, J.; Klionsky, D.J.; Tang, D. Posttranslational modification of autophagy-related proteins in macroautophagy. Autophagy 2015, 11, 28–45. [Google Scholar] [CrossRef] [Green Version]
  211. Fan, J.; Yang, X.; Li, J.; Shu, Z.; Dai, J.; Liu, X.; Li, B.; Jia, S.; Kou, X.; Yang, Y.; et al. Spermidine coupled with exercise rescues skeletal muscle atrophy from D-gal-induced aging rats through enhanced autophagy and reduced apoptosis via AMPK-FOXO3a signal pathway. Oncotarget 2017, 8, 17475–17490. [Google Scholar] [CrossRef] [PubMed]
  212. Mammucari, C.; Milan, G.; Romanello, V.; Masiero, E.; Rudolf, R.; Del Piccolo, P.; Burden, S.J.; Di Lisi, R.; Sandri, C.; Zhao, J.; et al. FoxO3 controls autophagy in skeletal muscle in vivo. Cell Metab. 2007, 6, 458–471. [Google Scholar] [CrossRef]
  213. Herrenbruck, A.R.; Bollinger, L.M. Role of skeletal muscle autophagy in high-fat-diet-induced obesity and exercise. Nutr. Rev. 2020, 78, 56–64. [Google Scholar] [CrossRef]
  214. Neel, B.A.; Lin, Y.; Pessin, J.E. Skeletal muscle autophagy: A new metabolic regulator. Trends Endocrinol. Metab. TEM 2013, 24, 635–643. [Google Scholar] [CrossRef]
  215. Wang, X.; Lin, Y.; Liu, S.; Zhu, Y.; Lu, K.; Broering, R.; Lu, M. O-GlcNAcylation modulates HBV replication through regulating cellular autophagy at multiple levels. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2020, 34, 14473–14489. [Google Scholar] [CrossRef]
  216. Shen, Y.; Czaja, M.J. A Novel Mechanism of Starvation-Stimulated Hepatic Autophagy: Calcium-Induced O-GlcNAc-Dependent Signaling. Hepatology 2019, 69, 446–448. [Google Scholar] [CrossRef]
  217. Kim, K.; Yoo, H.C.; Kim, B.G.; Kim, S.; Sung, Y.; Yoon, I.; Yu, Y.C.; Park, S.J.; Kim, J.H.; Myung, K.; et al. O-GlcNAc modification of leucyl-tRNA synthetase 1 integrates leucine and glucose availability to regulate mTORC1 and the metabolic fate of leucine. Nat. Commun. 2022, 13, 2904. [Google Scholar] [CrossRef] [PubMed]
  218. Wang, P.; Hanover, J.A. Nutrient-driven O-GlcNAc cycling influences autophagic flux and neurodegenerative proteotoxicity. Autophagy 2013, 9, 604–606. [Google Scholar] [CrossRef] [PubMed]
  219. Kumar, A.; Singh, P.K.; Parihar, R.; Dwivedi, V.; Lakhotia, S.C.; Ganesh, S. Decreased O-linked GlcNAcylation protects from cytotoxicity mediated by huntingtin exon1 protein fragment. J. Biol. Chem. 2014, 289, 13543–13553. [Google Scholar] [CrossRef] [PubMed]
  220. Cui, Z.; Scruggs, S.B.; Gilda, J.E.; Ping, P.; Gomes, A.V. Regulation of cardiac proteasomes by ubiquitination, SUMOylation, and beyond. J. Mol. Cell. Cardiol. 2014, 71, 32–42. [Google Scholar] [CrossRef] [Green Version]
  221. Zhang, F.; Hu, Y.; Huang, P.; Toleman, C.A.; Paterson, A.J.; Kudlow, J.E. Proteasome function is regulated by cyclic AMP-dependent protein kinase through phosphorylation of Rpt6. J. Biol. Chem. 2007, 282, 22460–22471. [Google Scholar] [CrossRef] [PubMed]
  222. Liu, H.; Yu, S.; Zhang, H.; Xu, J. Identification of nitric oxide as an endogenous inhibitor of 26S proteasomes in vascular endothelial cells. PLoS ONE 2014, 9, e98486. [Google Scholar] [CrossRef] [PubMed]
  223. Yan, W.; Cao, M.; Ruan, X.; Jiang, L.; Lee, S.; Lemanek, A.; Ghassemian, M.; Pizzo, D.P.; Wan, Y.; Qiao, Y.; et al. Cancer-cell-secreted miR-122 suppresses O-GlcNAcylation to promote skeletal muscle proteolysis. Nat. Cell Biol. 2022, 24, 793–804. [Google Scholar] [CrossRef] [PubMed]
  224. Wang, P.; Lazarus, B.D.; Forsythe, M.E.; Love, D.C.; Krause, M.W.; Hanover, J.A. O-GlcNAc cycling mutants modulate proteotoxicity in Caenorhabditis elegans models of human neurodegenerative diseases. Proc. Natl. Acad. Sci. USA 2012, 109, 17669–17674. [Google Scholar] [CrossRef]
  225. Love, D.C.; Ghosh, S.; Mondoux, M.A.; Fukushige, T.; Wang, P.; Wilson, M.A.; Iser, W.B.; Wolkow, C.A.; Krause, M.W.; Hanover, J.A. Dynamic O-GlcNAc cycling at promoters of Caenorhabditis elegans genes regulating longevity, stress, and immunity. Proc. Natl. Acad. Sci. USA 2010, 107, 7413–7418. [Google Scholar] [CrossRef]
  226. Park, S.; Lee, Y.; Pak, J.W.; Kim, H.; Choi, H.; Kim, J.W.; Roth, J.; Cho, J.W. O-GlcNAc modification is essential for the regulation of autophagy in Drosophila melanogaster. Cell. Mol. Life Sci. CMLS 2015, 72, 3173–3183. [Google Scholar] [CrossRef]
  227. Huang, L.; Yuan, P.; Yu, P.; Kong, Q.; Xu, Z.; Yan, X.; Shen, Y.; Yang, J.; Wan, R.; Hong, K.; et al. O-GlcNAc-modified SNAP29 inhibits autophagy-mediated degradation via the disturbed SNAP29-STX17-VAMP8 complex and exacerbates myocardial injury in type I diabetic rats. Int. J. Mol. Med. 2018, 42, 3278–3290. [Google Scholar] [CrossRef]
  228. Dodson, M.; Liu, P.; Jiang, T.; Ambrose, A.J.; Luo, G.; de la Vega, M.R.; Cholanians, A.B.; Wong, P.K.; Chapman, E.; Zhang, D.D. Increased O-GlcNAcylation of SNAP29 Drives Arsenic-Induced Autophagic Dysfunction. Mol. Cell. Biol. 2018, 38, e00595-17. [Google Scholar] [CrossRef]
  229. Zhang, X.; Wang, L.; Lak, B.; Li, J.; Jokitalo, E.; Wang, Y. GRASP55 Senses Glucose Deprivation through O-GlcNAcylation to Promote Autophagosome-Lysosome Fusion. Dev. Cell 2018, 45, 245–261.e246. [Google Scholar] [CrossRef]
  230. Shi, Y.; Yan, S.; Shao, G.C.; Wang, J.; Jian, Y.P.; Liu, B.; Yuan, Y.; Qin, K.; Nai, S.; Huang, X.; et al. O-GlcNAcylation stabilizes the autophagy-initiating kinase ULK1 by inhibiting chaperone-mediated autophagy upon HPV infection. J. Biol. Chem. 2022, 298, 102341. [Google Scholar] [CrossRef]
Figure 1. Protein O-GlcNAcylation homeostasis is maintained by its only pair of process enzymes, OGA and OGT. To date, thousands of proteins have been identified as O-GlcNAcylated, and these proteins are diverse including enzymes, transcription factors, signaling proteins, mitochondrial proteins, among others. O-GlcNAcylation homeostasis of these proteins is precisely and rapidly regulated only by OGT and OGA. Faced with mild stress and acute adverse stimulation, OGT and OGA form a “buffer” to make O-GlcNAcylation fluctuate adaptively within the “optimal zone”. Strong stress or chronic adverse stimulation beyond the buffer threshold of OGT and OGA results in O-GlcNAcylation levels exceeding the optimal zone. Both hyper- and hypo-O-GlcNAcylation beyond the “buffer region” lead to cellular dysfunction and other harmful effects.
Figure 1. Protein O-GlcNAcylation homeostasis is maintained by its only pair of process enzymes, OGA and OGT. To date, thousands of proteins have been identified as O-GlcNAcylated, and these proteins are diverse including enzymes, transcription factors, signaling proteins, mitochondrial proteins, among others. O-GlcNAcylation homeostasis of these proteins is precisely and rapidly regulated only by OGT and OGA. Faced with mild stress and acute adverse stimulation, OGT and OGA form a “buffer” to make O-GlcNAcylation fluctuate adaptively within the “optimal zone”. Strong stress or chronic adverse stimulation beyond the buffer threshold of OGT and OGA results in O-GlcNAcylation levels exceeding the optimal zone. Both hyper- and hypo-O-GlcNAcylation beyond the “buffer region” lead to cellular dysfunction and other harmful effects.
Metabolites 12 00888 g001
Figure 2. O-GlcNAcylation is linked by the HBP to nutrient availability and thus its regulation of the complex metabolic network. Two percent to 3% of intracellular glucose enters the HBP and undergoes a series of chemical reactions to produce the end product UDP-GlcNAc. The UDP-GlcNAc in the nucleus, cytoplasm and membrane is utilized for O-GlcNAcylation. UDP-GlcNAc makes the dynamic O-GlcNAcylation extremely sensitive to changes in cellular nutrients used as a sensor of the functional state of multiple pathways such as glycolysis, TCA cycle, pentose phosphate pathway (PPP), glycogen synthesis and catabolism, etc. Meanwhile, the interconversion of intermediates between the HBP, polyol pathway, fructose, PPP, glycogen, glycolysis and TCA cycle greatly enhanced the nutritional sensitivity of O-GlcNAcylation. A portion of UDP-GlcNAc is transported to the endoplasmic reticulum and the Golgi apparatus while being used for other O-glycosylation and N-glycosylation. Other fractions of UDP-GlcNAc are used for the biosynthesis of proteoglycans, hyaluronic acid, glycolipids, glycolphosphatidylinositol anchoring, etc.
Figure 2. O-GlcNAcylation is linked by the HBP to nutrient availability and thus its regulation of the complex metabolic network. Two percent to 3% of intracellular glucose enters the HBP and undergoes a series of chemical reactions to produce the end product UDP-GlcNAc. The UDP-GlcNAc in the nucleus, cytoplasm and membrane is utilized for O-GlcNAcylation. UDP-GlcNAc makes the dynamic O-GlcNAcylation extremely sensitive to changes in cellular nutrients used as a sensor of the functional state of multiple pathways such as glycolysis, TCA cycle, pentose phosphate pathway (PPP), glycogen synthesis and catabolism, etc. Meanwhile, the interconversion of intermediates between the HBP, polyol pathway, fructose, PPP, glycogen, glycolysis and TCA cycle greatly enhanced the nutritional sensitivity of O-GlcNAcylation. A portion of UDP-GlcNAc is transported to the endoplasmic reticulum and the Golgi apparatus while being used for other O-glycosylation and N-glycosylation. Other fractions of UDP-GlcNAc are used for the biosynthesis of proteoglycans, hyaluronic acid, glycolipids, glycolphosphatidylinositol anchoring, etc.
Metabolites 12 00888 g002
Figure 3. Spatiotemporal regulation of O-GlcNAcylation on insulin signal transduction of skeletal muscle. Skeletal muscle is the major insulin-sensitive tissue, responsible for about 75% of insulin-stimulated glucose uptake in the whole body, which is critical for glucose homeostasis and energy metabolism. The harmonious balance between O-GlcNAcylation and phosphorylation is necessary for the normal operation of insulin signaling under normal physiology, and the disruption of this homeostasis leads to the impairment of insulin-signaling cascade. O-GlcNAcylation has been described as protecting cells from acute stress but harmful to chronic and persistent stress in various organs or tissues. For example, almost all insulin-sensitive tissues have a hyper O-GlcNAcylation level and many related complications are under hyperglycemia. This adverse effect is achieved through impaired glucose utilization or glucose toxicity paradigms and may lead to the progression of a variety of diseases. Therefore, chronic elevation of O-GlcNAcylation is considered as a mechanism for the development of insulin resistance.
Figure 3. Spatiotemporal regulation of O-GlcNAcylation on insulin signal transduction of skeletal muscle. Skeletal muscle is the major insulin-sensitive tissue, responsible for about 75% of insulin-stimulated glucose uptake in the whole body, which is critical for glucose homeostasis and energy metabolism. The harmonious balance between O-GlcNAcylation and phosphorylation is necessary for the normal operation of insulin signaling under normal physiology, and the disruption of this homeostasis leads to the impairment of insulin-signaling cascade. O-GlcNAcylation has been described as protecting cells from acute stress but harmful to chronic and persistent stress in various organs or tissues. For example, almost all insulin-sensitive tissues have a hyper O-GlcNAcylation level and many related complications are under hyperglycemia. This adverse effect is achieved through impaired glucose utilization or glucose toxicity paradigms and may lead to the progression of a variety of diseases. Therefore, chronic elevation of O-GlcNAcylation is considered as a mechanism for the development of insulin resistance.
Metabolites 12 00888 g003
Figure 4. The main target of O-GlcNAcylation involved in autophagy process of skeletal muscle. O-GlcNAcylation is involved in almost all processes of autophagy including initiation, nucleation, expansion and fusion in addition to degradation. First, there is extremely extensive crosstalk between O-GlcNAcylation and autophagic signal transduction. Activation of AMPK in response to energy deprivation inhibits mTOR and activates ULK1. Acute activation of AMPK alters the substrate selectivity and nuclear localization of OGT in myotubes, which is closely related to the phosphorylation of OGT at Thr444 by AMPK. In contrast, OGT catalyzes the O-GlcNAcylation of α and γ subunits of AMPK, and these modifications increase with increasing AMPK activity. There is crosstalk between O-GlcNAcylation and mTOR signaling including their mutual regulation of each other’s stability. Meanwhile, OGT regulates autophagy gene expression and autophagy structure through AKT and FoxO3. At the initiation of autophagy, ULK1-ATG13-ATG101-FIP200 complex and vacuolar protein sorting (VPS34)-Beclin 1 complex coordinate to drive the formation of isolation membrane, which is inhibited by mTORC1 and inactivated by Bcl-2 inactivates, respectively. O-GlcNAcylation of ULK1 at Thr754 is essential for ATG14L binding and phosphorylation, allowing the activation of VPS34 and leading to the production of a phosphatidylinositol 3-phosphate cloud locally. Beclin-1 and Bcl-2 are targets for O-GlcNAcylation. OGT deletion was associated with increased ATG expression and autophagy lysosomes. Inhibition of OGT significantly increased the level of LC3-II and LC3 punctate structure, decreased the level of autophagy substrate SQSTM1, and increased the expression of lysosome-associated membrane protein (LAMP) 1 and 2. These evidences are sufficient to indicate that O-GlcNAcylatio plays a pivotal role in regulating the expansion of autophagy. Subsequently, O-GlcNAcylation sites of SNAP-29 at Ser2, Ser61, Thr130 and Ser153 disturbed the SNAP29-syntaxin-17 (STX17)-vesicle-associated membrane protein 8 (VAMP8) complex, which regulates autophagosome maturation.
Figure 4. The main target of O-GlcNAcylation involved in autophagy process of skeletal muscle. O-GlcNAcylation is involved in almost all processes of autophagy including initiation, nucleation, expansion and fusion in addition to degradation. First, there is extremely extensive crosstalk between O-GlcNAcylation and autophagic signal transduction. Activation of AMPK in response to energy deprivation inhibits mTOR and activates ULK1. Acute activation of AMPK alters the substrate selectivity and nuclear localization of OGT in myotubes, which is closely related to the phosphorylation of OGT at Thr444 by AMPK. In contrast, OGT catalyzes the O-GlcNAcylation of α and γ subunits of AMPK, and these modifications increase with increasing AMPK activity. There is crosstalk between O-GlcNAcylation and mTOR signaling including their mutual regulation of each other’s stability. Meanwhile, OGT regulates autophagy gene expression and autophagy structure through AKT and FoxO3. At the initiation of autophagy, ULK1-ATG13-ATG101-FIP200 complex and vacuolar protein sorting (VPS34)-Beclin 1 complex coordinate to drive the formation of isolation membrane, which is inhibited by mTORC1 and inactivated by Bcl-2 inactivates, respectively. O-GlcNAcylation of ULK1 at Thr754 is essential for ATG14L binding and phosphorylation, allowing the activation of VPS34 and leading to the production of a phosphatidylinositol 3-phosphate cloud locally. Beclin-1 and Bcl-2 are targets for O-GlcNAcylation. OGT deletion was associated with increased ATG expression and autophagy lysosomes. Inhibition of OGT significantly increased the level of LC3-II and LC3 punctate structure, decreased the level of autophagy substrate SQSTM1, and increased the expression of lysosome-associated membrane protein (LAMP) 1 and 2. These evidences are sufficient to indicate that O-GlcNAcylatio plays a pivotal role in regulating the expansion of autophagy. Subsequently, O-GlcNAcylation sites of SNAP-29 at Ser2, Ser61, Thr130 and Ser153 disturbed the SNAP29-syntaxin-17 (STX17)-vesicle-associated membrane protein 8 (VAMP8) complex, which regulates autophagosome maturation.
Metabolites 12 00888 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Y.; Hu, Y.; Li, S. Protein O-GlcNAcylation in Metabolic Modulation of Skeletal Muscle: A Bright but Long Way to Go. Metabolites 2022, 12, 888. https://doi.org/10.3390/metabo12100888

AMA Style

Liu Y, Hu Y, Li S. Protein O-GlcNAcylation in Metabolic Modulation of Skeletal Muscle: A Bright but Long Way to Go. Metabolites. 2022; 12(10):888. https://doi.org/10.3390/metabo12100888

Chicago/Turabian Style

Liu, Yang, Yajie Hu, and Shize Li. 2022. "Protein O-GlcNAcylation in Metabolic Modulation of Skeletal Muscle: A Bright but Long Way to Go" Metabolites 12, no. 10: 888. https://doi.org/10.3390/metabo12100888

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop