Next Article in Journal
Antioxidant Activity Analysis of Native Actinidia arguta Cultivars
Next Article in Special Issue
The Structure of the Cardiac Mitochondria Respirasome Is Adapted for the β-Oxidation of Fatty Acids
Previous Article in Journal
Preparative and Catalytic Properties of MoVI Mononuclear and Metallosupramolecular Coordination Assemblies Bearing Hydrazonato Ligands
Previous Article in Special Issue
The Research Progress of Mitochondrial Transplantation in the Treatment of Mitochondrial Defective Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial Dysfunction, Oxidative Stress, and Inter-Organ Miscommunications in T2D Progression

by
Rajakrishnan Veluthakal
*,
Diana Esparza
,
Joseph M. Hoolachan
,
Rekha Balakrishnan
,
Miwon Ahn
,
Eunjin Oh
,
Chathurani S. Jayasena
and
Debbie C. Thurmond
*
Department of Molecular and Cellular Endocrinology, Arthur Riggs Diabetes & Metabolism Research Institute, City of Hope Beckman Research Institute, 1500 E. Duarte Rd, Duarte, CA 91010, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(3), 1504; https://doi.org/10.3390/ijms25031504
Submission received: 22 December 2023 / Revised: 15 January 2024 / Accepted: 23 January 2024 / Published: 25 January 2024
(This article belongs to the Special Issue Mitochondria in Human Health and Disease 2.0)

Abstract

:
Type 2 diabetes (T2D) is a heterogenous disease, and conventionally, peripheral insulin resistance (IR) was thought to precede islet β-cell dysfunction, promoting progression from prediabetes to T2D. New evidence suggests that T2D-lean individuals experience early β-cell dysfunction without significant IR. Regardless of the primary event (i.e., IR vs. β-cell dysfunction) that contributes to dysglycemia, significant early-onset oxidative damage and mitochondrial dysfunction in multiple metabolic tissues may be a driver of T2D onset and progression. Oxidative stress, defined as the generation of reactive oxygen species (ROS), is mediated by hyperglycemia alone or in combination with lipids. Physiological oxidative stress promotes inter-tissue communication, while pathological oxidative stress promotes inter-tissue mis-communication, and new evidence suggests that this is mediated via extracellular vesicles (EVs), including mitochondria containing EVs. Under metabolic-related stress conditions, EV-mediated cross-talk between β-cells and skeletal muscle likely trigger mitochondrial anomalies leading to prediabetes and T2D. This article reviews the underlying molecular mechanisms in ROS-related pathogenesis of prediabetes, including mitophagy and mitochondrial dynamics due to oxidative stress. Further, this review will describe the potential of various therapeutic avenues for attenuating oxidative damage, reversing prediabetes and preventing progression to T2D.

1. Introduction

1.1. Prediabetes and Type 2 Diabetes

Over the last decade, type 2 diabetes (T2D) has emerged as a complex multifactorial heterogenous disease [1]. About 90–95% of all diabetes cases worldwide are composed of T2D cases [2,3]. Individuals are clinically diagnosed with T2D when there is a relative insulin deficiency (i.e., islet β-cell dysfunction) combined with peripheral insulin resistance (IR) [3]. Alarmingly, 541 million adults worldwide are currently living with prediabetes, a transitional asymptomatic hyperglycemic state between normoglycemia and T2D, and these individuals are at a significant risk of developing T2D and other metabolic complications that impact their lifespan and quality of life [2]. Four T2D archetype profiles in newly diagnosed individuals were recently identified based on differences in disease progression, genetic risk scores, and circulating omics biomarkers (Figure 1) and they were: (1) lean and insulin deficient, (2) obese and insulin sensitive, (3) obese and IR, and (4) a severe phenotype with obesity, IR, dyslipidemia, and impaired β-cell glucose sensitivity that progresses rapidly to frank T2D that required the most anti-diabetic interventions [4].
IR is a hallmark of prediabetes and is characterized by an impaired cellular response to insulin stimulation at the peripheral tissues, such as skeletal muscle and adipose tissue [3]. This leads to dysglycemia (i.e., impaired fasting glucose and/or glucose tolerance) and metabolic dyshomeostasis. Conventionally, skeletal muscle IR is thought to precede islet β-cell dysfunction and be the primary event driving dysglycemia toward T2D. While this may be the case in obesity-related T2D, this notion is challenged by the observation that various populations have varying degrees of insulin secretory capacity and IR [5]. Critically, there is a high incidence of prediabetes and T2D in lean non-white populations with a BMI index <25 kg/m2 and these individuals present with insulin deficiency, rather than IR [6,7,8,9]. It is noteworthy however that the BMI may not realistically capture adiposity.
Regardless of the primary event (i.e., IR vs. β-cell dysfunction) driving dysglycemia, compelling mounting evidence suggests that prolonged pathological oxidative stress within metabolic tissues (e.g., skeletal muscle, adipose tissue, pancreatic islet β-cells) may be a crucial contributor to the metabolic abnormalities present during prediabetes and drive progression to T2D [10,11,12,13,14,15,16,17]. Thus, gaining a deeper understanding of the mechanisms that drive pathological oxidative stress could aid in the development of novel therapeutic strategies that halt/prevent progression to T2D.

1.2. Oxidative Stress

Oxidative stress is defined as an impairment of the defensive antioxidant systems and an elevation of the reactive oxygen species (ROS), which include superoxide anion radical (O2•−), hydrogen peroxide (H2O2) and hydroxyl radical (OH), as well as reactive nitrogen radicals (RNS) and non-radical peroxynitrite (ONOO) [18]. Compartmentalized and controlled ROS production has emerged as a physiological mediator of many cellular responses [19,20], including glucose-stimulated insulin secretion (GSIS) by islet β-cells and glucose-uptake related to skeletal muscle function [21,22,23,24,25].
Mitochondrial ROS (mROS) and extramitochondrial sources contribute to ROS production and there is the potential for these various sources to interact with one another to alter redox balance and promote oxidative stress [26,27]. It is well-known that glycolytic and oxidative metabolism of glucose, as well as fatty acid oxidation leads to adenosine triphosphate (ATP) generation in the mitochondria [28,29,30]. The major source of ROS generation is mitochondrial respiration during ATP production; electrons that escape the electron transport chain (ETC) reacts with oxygen to produce O2•− and H2O2 (reviewed in [31]). Nicotinamide adenosine dinucleotide phosphate oxidases (NOXs), a membrane-associated protein complex, also promotes the generation of O2•− using oxidation of nicotinamide adenosine dinucleotide phosphate (NADPH). In addition to the expression and localization to the cell membrane, NOX proteins also reside in the endoplasmic reticulum (ER) and in the mitochondria [32,33,34]. Peroxisomes that are armed with oxidases constitute another source of H2O2 production and are essential for lipid metabolism.
ROS is a powerful oxidant that damages many biological molecules [35,36] (Figure 2). In addition to being a significant source of cellular ROS, mitochondria are targets of cellular ROS damage. Mitochondrial DNA (mtDNA), protein, and membrane associated lipids are susceptible to oxidative stress leading to mitochondrial dysfunction [37]. Mitochondrial dysfunction is characterized by diminished ATP production, altered mitochondrial enzymatic activity, impaired redox balance and excessive generation of ROS. Thus, mitochondrial dysfunction has the capacity to further enhance oxidative stress [38,39].
Strict maintenance of glycemic control is necessary for maintaining mitochondrial function and morphology in humans and animals [40,41,42]. Mitochondrial metabolism is reduced in obese-IR individuals [43]. Furthermore, T2D or IR individuals show a deficiency for the mitochondrial biogenesis master regulator, peroxisome proliferator-activated receptor coactivator-1 alpha (PGC1α) in their skeletal muscle [44,45]. Plasma ROS production was also recently found to be significantly elevated in individuals with prediabetes and those with newly diagnosed T2D [46]. However, transient ROS elevations in skeletal muscle and islet β-cells are thought to be beneficial, adaptive and necessary in health. The threshold for ROS elevations and associated mitochondrial dysfunction that is considered detrimental to the specialized functions of skeletal muscle and pancreatic islet β-cells that lead to metabolic dyshomeostasis during prediabetes and progression to T2D remains to be established. This review will highlight new findings related to the underlying molecular mechanisms in ROS-related pathogenesis of (pre)diabetes, including mitochondrial dysfunction and inter-organ miscommunication that impacts mitochondrial function due to pathological oxidative stress. We have also attempted to organize information in this review as a continuum and present a broader picture on the contributory factors that promote T2D development including IR and β-cell dysfunction. Further, this review will describe the potential of therapeutic avenues for attenuating oxidative damage, reversing prediabetes and preventing progression to T2D.

2. Physiological ROS Production and Glucose Metabolism in Health

There is emerging appreciation that transient bursts of ROS from mitochondrial and extramitochondrial sources facilitates insulin release in response to glucose from islet β-cells and promotes skeletal muscle glucose uptake in health. Gaining an understanding of the various cellular ROS sources that contribute to glucose metabolism in health is essential because the dysregulation of any of these cellular ROS pathways has the potential to promote mitochondrial dysfunction and progression to T2D.

2.1. Glucose Sensing, Insulin Secretion and Islet β-Cells

Insulin is a hormone that is essential for whole-body glucose homeostasis and is released by islet β-cells in response to glucose. ROS is emerging as a signaling component necessary for islet β-cell glucose sensing and insulin secretion (Figure 3). The presence of glucose stimulates H2O2 generation [22,47] and has been shown to alter the intracellular redox status [22]. Further, an increase in intracellular H2O2 leads to an increase in insulin secretion [21]. Suppression of glucose-induced H2O2 accumulation by antioxidants hinders GSIS [21], and ROS activates the ryanodine receptor, which subsequently provides the necessary increase in intracellular Ca2+ for insulin release [48]. Islet β-cells express cell membrane-localized NOX2 [49], and the generation of ROS by guanine nucleotide exchange factor cytohesin-2 (ARNO) signaling promotes NOX2 complex assembly and GSIS in rat INS 832/13 clonal β-cells [50].
Mitochondrial blockers were shown to enhance mROS levels and elevate insulin release similar to levels stimulated by glucose [51]. Knockdown of the mitochondrial uncoupling protein 2 (UCP2) increased GSIS from INS-1E cells [52], and this was ablated by exposure to the antioxidant MnTMPyP [53]. NOX4 is reported to reside within the mitochondria and is able to generate both O2•− and H2O2 [54]. β-cell–specific NOX4 knockout mice (NOX4βKO) and from NOX4-silenced or catalase-overexpressing INS-1E clonal β-cell studies showed that the first phase of insulin secretion was completely abolished [22]. Furthermore, lentiviral NOX4 overexpression or enhancing H2O2 rescued GSIS from NOX4βKO mice [22]. NOX4βKO mice exhibited impaired glucose tolerance and peripheral IR [22], thus, demonstrating the in vivo role of NOX4 derived ROS in insulin release. In contrast, using a pharmacological NOX4 inhibitor (GLX7013114) in human islets and human EndoC-βH1 cells showed maximal oxygen consumption rates were increased in islets after acute NOX4 inhibition [55]. In EndoC-βH1 cells, NOX4 inhibition increased the mitochondrial membrane potential, mROS and the ATP/ADP ratio. Furthermore, the insulin release from EndoC-βH1 cells at a high glucose concentration increased with NOX4 inhibition, thus suggesting that pharmacological alleviation of NOX4 could result in enhanced insulin release. However, the discrepancies in the studies may be due to small changes in ROS levels at different microenvironments that escape detection, unless a specific probe is utilized to explicitly target a particular microenvironment to study the regulation of ROS from that compartment.
In the past several decades, the islet biology field has looked at conventional antioxidant systems to protect islet β-cells from oxidative stress. New emerging evidence suggests that there are several other alternate antioxidant systems conferring protection to β-cells, such as the peroxiredoxins (PRX), thioredoxins (TRX), thioredoxin reductases (TRXR), and NRF2 [56,57]. NRF2 is implicated as a master regulator of several antioxidant genes based on its ability to confer β-cell protection against experimentally-induced supraphysiological concentrations of ROS [58]. The varying degree with which these different antioxidants offer protection against oxidative stress may depend on the generation of ROS at different subcellular microdomains and thus escape detection due to an inability to measure such changes due to their highly reactive nature.

2.2. Glucose Uptake and Skeletal Muscle Function

In the post-prandial state in healthy individuals, skeletal muscle accounts for ~70–90% of glucose uptake from the blood [59,60]. Transient levels of ROS produced in a controlled fashion play crucial roles in skeletal muscle cells, including the control of gene expression, regulation of cell signaling pathways, and modulation of skeletal muscle force production (reviewed in [61]). For example: (i) rat extensor digitorum longus muscles increased glucose uptake when exposed to transient elevated oxidative stress, and catalase exposure blunted glucose uptake [62]; (ii) H2O2-stimulated glucose uptake was completely inhibited by the phosphatidylinositol 3-kinase (PI3K) inhibitor wortmannin, but not the nitric oxide inhibitor NG-monomethyl-l-arginine [62]; (iii) ROS and NO have been implicated in skeletal muscle glucose uptake, via an AMP-activated protein kinase (AMPK) independent mechanism during contraction [23]. Moreover, NOX2-derived cytosolic ROS is required for insulin-stimulated glucose transporter type 4 (GLUT4)-mediated glucose uptake in skeletal muscle [24]. Relatedly, exercise-induced GLUT4 translocation and glucose uptake are impaired in knockout mouse models of NOX2 assembly components, Rac1 and p47phox [24]. The requirement for skeletal muscle NOX4 in adaptive responses that promote insulin sensitivity and maintain redox balance was recently reported [25]. Further, stabilization of the NFE2L2-mediated antioxidant defense by NOX4-derived H2O2 was necessary for maintaining insulin sensitivity in a cell autonomous manner [25]. Collectively, these findings support the notion that transient levels of ROS are important for skeletal muscle function.

3. Impaired Mitochondrial Dynamics and Mitophagy in Skeletal Muscle Insulin Resistance

The mitochondrial networks across the skeletal muscle fiber types undergo dynamic and heterogeneous alterations and are vital components for muscle plasticity [63] that permit adaptations to the fluctuations in skeletal muscle energy demands [64]. The progressive decline in skeletal muscle insulin sensitivity, culminating in chronic IR, is a major contributor to T2D onset. Abnormal skeletal muscle mROS levels are associated with high-fat and high-glucose diet-induced IR in humans and in preclinical models [25,65,66].

3.1. Impaired Mitochondrial Dynamics in Skeletal Muscle IR

Mitochondrial dynamics consists of the coordinated cycling between mitochondrial fission (division of one mitochondrion into two daughter mitochondria) and fusion (the union of two mitochondria into one mitochondrion). Typically, mitochondrial fragmentation phenotypes due to elevated mitochondrial fission activity are linked to mitochondrial dysfunction and are observed during elevated stress levels and cell death (reviewed in [67]) [68,69]. Clinical evidence from individuals with T2D supports the concept that progressive IR is associated with elevated levels of fragmented mitochondrial networks and diminished oxidative capacity [70].
A high-fat diet (HFD) increases the risk of skeletal muscle IR via increased intramyocellular lipid accumulation [64] increased ROS production [71] mitochondrial damage [72], muscle inflammation [73,74] and dysregulation of insulin signaling [75]. HFD-fed mice [76], as well as palmitate-treated human and rodent myotubes [77,78], showed increased mitochondrial fragmentation driven by the activation of the fission marker dynamin-related protein 1 (Drp1). HFD and/or palmitate can elevate ROS production via the mitochondrial ETC and facilitate the assembly of NOX by phosphorylating p47phox [79]. Furthermore, ROS-induced damaged mitochondria are marked for clearance via mitophagy [80]. In parallel, palmitate contributes toward IR by downregulating canonical AKT signaling [81]. Elevated mitochondrial fission is a preceding event to mitophagy [82,83]. Thus, targeting mitophagy has been speculated as a means to alleviate skeletal muscle IR.

3.2. Molecular Pathways Involved in Mitophagy: A Novel Role for STX4?

One of the early responses to oxidative stress is the induction of mitophagy, a specialized form of autophagy that selectively disposes of aging, damaged or dysfunctional mitochondria (reviewed in [84]). Below we detail the canonical and non-canonical mitophagy pathways (Figure 4) [85,86,87,88,89,90,91], and integrate the emerging novel putative role for the exocytosis protein, Syntaxin 4 (STX4) in skeletal muscle IR-related mitophagy [76].
Canonical PARKIN-dependent mitophagy: The most defined mechanism of mitophagy is the ubiquitin-dependent phosphatase and tense homologue (PTEN)-induced putative kinase 1 (PINK1)-Parkin RBR E3 ubiquitin-protein ligase (PARKIN) pathway (Figure 4a). In healthy mitochondria, the cytosolic PINK1 serine/threonine kinase that translocates towards the mitochondria is imported across the outer- and inner-mitochondrial membranes (OMM and IMM) through their respective translocases (translocase of the outer membrane [TOM] and translocase of the inner membrane [TIM] complexes) prior to cleavage by presenilin-associated rhomboid like (PARL) proteases [92,93], resulting in a low basal level of PINK1 around healthy mitochondria. However, upon mitochondrial damage, the decreased membrane potential stabilizes OMM-localized PINK1, triggering autophosphorylation of Ser-228 and Ser-402 sites for PINK1 activation [89,94]. Activated OMM-localized PINK1 promotes PARKIN recruitment to the damaged mitochondria via phosphorylation of Ser-65 on OMM protein-associated ubiquitin [86,87,88,95]. Ser-65 phosphorylation of the recruited PARKIN activates its E3-ubiquitin ligase activity triggering a feedback loop of further Ser-65 phosphorylation on poly-ubiquitinated chains of OMM proteins, enabling more translocation of PARKIN proteins towards the mitochondria [86,87,88,95]. This sequence culminates in the recruitment of LC3-positive autophagosomes via autophagy receptors including p62 [85] and optineurin [96,97]).
Non-canonical receptor-dependent mitophagy: On the other hand, the receptor-dependent form of mitophagy does not rely on Ser-65 phosphorylated polyubiquitinated chains. Instead, OMM- and IMM-anchored proteins in possession of LC3-interacting motifs recruit LC3-positive autophagosomes by direct interaction (Figure 4b) [98,99,100]. To date, the most defined examples studied in mammalian cells include BCL2/adenovirus E1B 19 kDa protein-interacting protein 3-like (BNIP3L/NIX) [90,101] and FUN14 domain containing 1 (FUNDC1) [91,98,102,103] and most recently prohibitin-2, an IMM protein with a LC3-interacting motif that is involved in both ubiquitin- and receptor-dependent mitophagy [100,104].
Is STX4 a novel mitophagy associated protein?: STX4, is classically known as a key regulator of insulin-dependent glucose uptake by skeletal muscle, via its role as a SNARE protein at the sarcolemmal and t-tubule membrane surfaces [105,106]. We recently highlighted a novel localization for STX4 at the OMM, as well as a distinct mitochondrial role for STX4 at that locale [76]. Skeletal muscle-specific STX4 enrichment in HFD-fed mice led to reversal of IR, enhanced spontaneous activity and respiratory exchange ratio, as well as improvements in mitochondrial respiration and dynamics [76]. STX4 was found in the skeletal muscle mitochondrial fraction, and STX4 enrichment reduced skeletal muscle mitochondrial fragmentation and deactivated Drp1 in an AMPK-dependent manner [76]. Although this study focused on the role of STX4 in mitochondrial dynamics, studies have shown that mitochondrial fission proteins are also known regulators of mitophagy. For example, Drp1 within the mammalian brain, heart and pluripotent stem cells [107,108], and most recently the IMM fission protein MTFP1, have been identified as being involved in both PINK1-PARKIN and receptor-dependent mitophagy in human oral cancer cell lines [109].
Taken together, future studies focused on mitochondrial membrane proteins such as the recently discovered OMM-associated STX4 [76] may reveal a new player in mitophagy as it remains to be determined whether the absence of fragmented mitochondria in both lean and HFD skeletal muscle in STX4-enriched mice was due to increased repair and/or mitophagy for their clearance [76].

3.3. Mitophagy in Skeletal Muscle IR: Helpful or Harmful?

Although mitophagy is a key regulatory process for the maintenance of cellular homeostasis, in the diabetes field, a debate lingers on whether targeting mitophagy is beneficial for treating IR in prediabetes and T2D. Most pre-clinical in vivo studies focused on HFD-induced IR have evaluated PINK1/PARKIN-dependent mitophagy. For short-term HFD studies of 1–2 weeks [110,111] in mice, acute IR development corresponded with PINK1/PARKIN elevations in the skeletal muscle. In contrast, long-term HFD for 12–24 weeks in rodents [111,112] or 2 years in Rhesus monkeys [113] presented reduced expression of PINK1/PARKIN in the skeletal muscle. Despite the associative patterns of mitophagy with stages of IR across the pre-clinical models, these examples are limited by not distinguishing whether they represent the clinical characteristics of prediabetes or T2D. Nevertheless, one clinical study that focused on peripheral blood mononuclear cells reported that initial elevations of PARKIN/PINK1 and NIX mitophagy markers in prediabetes patients eventually decreased across newly diagnosed and advanced duration T2D patients [114]. Even though this study was not performed in insulin-sensitive tissues, the reported reduction of PINK1 levels in obese T2D skeletal muscle and progressive downregulation of mitochondrial protein in skeletal muscle from non-diabetes, to prediabetes and then T2D patients [115], could support a similar mitophagy model for diabetes. In this model, it is suggested that elevations of mitophagy associated with prediabetes and acute IR stages actually serve a protective role of mitigating accumulating mitochondrial damage (Figure 4c). However, as IR worsens towards T2D, mitophagy activity is impaired leading to excessive mROS production. Whether the decline of mitophagy is a cause or consequence of prediabetes progression into T2D remains to be determined. Although insulin-sensitizing lifestyle interventions of exercise and dietary restrictions [116] have been associated with recovering PINK1 and PARKIN in HFD-exposed IR models, emerging pre-clinical studies with flavonoids such as Bavachin and corylifol A in db/db mice [117], and isoflavonoid puerarin in palmitate-exposed L6 myotubes [118] also demonstrated improved insulin sensitivity with restored PARKIN/PINK1 activity. Although further therapeutic studies in the context of prediabetes and T2D are required, the evidence so far suggests that boosting mitophagy in the prediabetes stage could protect against accumulating mitochondrial damage and may serve as a useful preventative treatment.
Although evidence suggests that non-canonical receptor-mediated mitophagy may play a role in diabetes, the results have been quite mixed. Global FUNDC1-KO mice exhibited worsened HFD-induced IR [102]. Counterintuitively, skeletal muscle-specific FUNDC1-KO mice do not show obesity or IR when exposed to a 3-month-HFD [103]. Intriguingly, this was a result of skeletal muscle derived fibroblast growth factor 21 (FGF21) triggering thermogenesis in intercalating white adipose tissue via skeletal muscle-adipose cross talk [103]. Further, pancreatic-specific FUNDC1 overexpression also protected mice from a 4-month-HFD [91], suggesting that the requirement for FUNDC1 in modulating mitochondrial activity in response to IR may be tissue-dependent. A 3–12-week-HFD was associated with elevated BNIP3L/NIX in rats [119], whilst a later study investigating BNIP3L/NIX in C2C12 and primary human myotubes linked increased BNIP3L expression with impaired insulin signaling via mechanistic target of rapamycin kinase (mTOR)-dependent inhibition of insulin receptor substrate 1 (IRS-1) [101]. To date, no study has directly compared the differences between the PARKIN-dependent and -independent mitophagy pathways and their beneficial effects in prediabetes and T2D, and future studies are required to address this gap. PARKIN-independent mitophagy in T2D may be a tissue-specific requirement, and thus, an alternate model to the ubiquitin-dependent mitophagy model (Figure 4c).
Finally, it should be acknowledged that most of these mitophagy studies have been investigated under the obese and IR model of T2D (Figure 1). Interestingly, a recent 2023 study that measured PINK1 mRNA levels in the whole blood samples of obese-IR versus lean-T2D patients reported that decreased PINK1 was associated only with the obese cohort [120], suggesting that mitophagy may be different across the heterogenous T2D groups. Thus, the currently proposed mitophagy model (Figure 4c) may not be universally applicable to all T2D cohorts (Figure 1), and future studies need to investigate the impact of mitophagy in lean-T2D patients.

4. ROS-Related Molecular Pathways Underlying IR and T2D

T2D individuals show a deficit in skeletal muscle for the serine/threonine p21 (Cdc42/Rac1) activated kinase 1 (PAK1) [121], which was recently implicated in mitochondrial and glucose dyshomeostasis [121]. Muscle deficiency for PAK1 showed reduced expression of PGC1α [122], consistent with an essential role for PAK1 in maintaining redox balance via electron transport chain proteins in β-cells [123]. Several lines of evidence suggest that mROS-induced oxidative stress in skeletal muscle IR and T2D leads to maladaptive enhanced activation of the polyol pathway flux and advanced glycated end products (AGEs) and hexosamine pathways [124,125] (Figure 5). Further, hyperglycemia enhances polyol and hexosamine pathways leading to increased endogenous AGEs production (reviewed in [126]).
Polyol pathway: Chronic hyperglycemia in prediabetes and T2D induces the aldose reductase (AR)-dependent polyol pathway, which increases O2•− generation and alters glucose and lipid metabolism thereby contributing to mitochondrial damage. Typically, AR reduces the toxic aldehydes to alcohol, but in a chronically hyperglycemic environment, it reduces glucose to sorbitol at the expense of NADPH. Additionally, in T2D, the hexokinase pathway becomes saturated, and the decrease in cytosolic free NAD+/NADH ratio changes the redox potential and enhances diacylglycerol (DAG) and triacylglycerol synthesis [127,128,129]. The overall reduction in NADPH levels reduces the regeneration of reduced glutathione (GSH) from its oxidized form. Further modeling this, the skeletal muscle of young male Wistar rats fed a HFD for 14 days showed reduced glutathione (GSH) and disulfide (GSSG) glutathione levels in the soleus muscle, as compared with chow-fed controls [130]. Further, MKR mice show reduced physical capacity, strength, and muscle mass leading to diabetic myopathy and increasing the risk of liver disease due to polyol accumulation [124]. In a recent clinical trial testing the effect of oral glutathione (GSH) supplementation for 3 weeks in obese subjects with and without T2D, the whole-body insulin sensitivity and skeletal muscle GSH significantly increased in the GSH group; oddly however, no changes were seen in skeletal muscle mitochondrial H2O2 emission rate, a measure of oxidative stress, pre- and post-GSH supplementation [131]. Thus, while it was concluded that oral GSH supplementation improved insulin sensitivity in obese patients without and with T2D, the link to skeletal muscle oxidative stress remained unclear.
Hexosamine pathway: IR-induced elevated fatty acid oxidation enhances fructose 6-phosphate levels that in turn activates hexosamine pathway and contributes to skeletal muscle IR. The rate limiting enzyme of this pathway is glutamine: fructose 6-phosphate amidotransferase (GFAT), which converts fructose 6-phosphate to glucosamine 6-phosphate and glutamate; upon further metabolism, this hexosamine biosynthesis pathway yields essential substrates for glycosylation of proteins and lipids. Additional major products of this pathway that accumulate in cells include UDP-N-Acetylglucosamine (UDP-GlcNAc) and UDP-N-acetylgalactosamine (UDP-GalNAc). Importantly, increased flux through the hexosamine pathway impairs glucose transport, contributing to the development of muscle IR [132,133]. Hexosamine pathway activity was reported to cause insulin resistance in C57BL/6NJ mice fed with HFD, and was related to molecular changes such as plasma membrane cholesterol accumulation and loss of filamentous actin, which is required for GLUT4 vesicle translocation and glucose uptake [134,135,136]. GLUT4 is N-glycosylated, and the absence of this modification prevents its subcellular localization during insulin stimulation and elevated hexosamine pathway flux could affect the N-glycan branching of glucose transporters and affect glucose uptake [137]. Moreover, individuals with T2D were found to have high GFAT1 activity determined from skeletal muscle biopsies, and GFAT activity in patients is related to postprandial hyperglycemia, oxidative stress, and other diabetic complications [138,139].
AGEs: Mitochondrial dysfunction is increased by AGEs [140]. AGEs are lethal metabolites that elicit oxidative stress via the receptor for advanced glycation end products (RAGE). RAGE activation is reported in high levels in chronic hyperglycemia [141,142,143,144]. Intracellular accumulation of AGEs alters protein structures and elicits oxidative stress. AGEs were initially implicated in skeletal muscle dysfunction leading to gradual and generalized loss of muscle strength [145]. More recently, AGEs are highly associated with sarcopenia in T2D individuals, resulting in decreased grip strength, knee extension strength and breakdown of muscular proteins [146]. In rats, AGE-albumin administration induced IR decreased GLUT4 expression and increased levels of nuclear factor NF-κB p50 subunit (NF-κB1) in skeletal muscle. Taken together, agents that reduce the development of AGEs would be predicted to prevent mitochondrial dysfunction and oxidative stress-induced IR. However, AGEs are broadly expressed and are a muscle-selective-drug targeting challenge. Nevertheless, AGEs are under scrutiny as biomarkers for earlier detection of diabetes-risk in bio-accessible fluids.

5. Mitochondrial Dysfunction and Islet β-Cell Dysfunction/Failure

Chronic overnutrition leads to a hyperactive metabolic state in β-cells, and the subsequent metabolic stress is related to suprathreshold ROS levels in β-cells. High glucose (glucotoxicity) and lipids (lipotoxicity) are known to cause β-cell damage via ROS, resulting in impaired GSIS (reviewed in [147]), mitochondrial defects and epigenetic modifications [148]. Reduced mtDNA content in β-cells is a marker of oxidative stress [149]. Beta cells are particularly susceptible to oxidative stress due to their high metabolic rate combined with low levels of antioxidant enzymes [150,151]. Thus, any impairment in the pro-oxidant/anti-oxidant balance ultimately leads to the development of T2D (reviewed in [152]). A thioredoxin/reductase-dependent mechanism enables β-cells to acquire the capacity to combat micromolar levels of H2O2 (50 µM), whereas a 100 µM bolus dose of H2O2 reduces β-cell viability and induces DNA damage, suggesting that the threshold limit for β-cells to handle the pro-oxidant capacity is less than 100 µM [57,153]. It is appropriate to point out that lean normoglycemic progeny from T2D patients have increased serum levels of 8-OHdG, an oxidative stress marker [154]. This begs the question whether oxidative stress is present in genetically predisposed subjects and induces the β-cell dysfunction in lean subjects in the absence of IR. Additionally, a number of chromosome 21q genes that contribute to T2D in lean individuals were identified and KCNJ15 (potassium inwardly-rectifying channel, subfamily J), a known T2D susceptibility gene, was among them [155]. T2D human islets and in vitro exposure of high glucose demonstrated elevated levels of KCNJ15; whether this elevation is a cause or consequence of T2D is yet to be determined.

5.1. Abnormal Mitochondrial Function in PAK1-Deficient β-Cells

The molecular regulators of mitochondrial function in β-cells are just beginning to be elucidated. Islets from T2D donors show an 80% deficiency for PAK1 compared to control individuals, suggesting that PAK1 deficiency might contribute to T2D pathogenesis [156,157]. We recently reported a vital requirement for PAK1 in modulating ROS in β-cells via its interaction with mitochondrial NADH:ubiquinone oxidoreductase subunit A12 (NDUFA12) [123], a protein that is critically required for oxidative stress regulation [158] (Figure 3). Beta cells deficient for PAK1 exhibited decreased mitochondrial ETC proteins (CI, CII and CIV), impaired mitochondrial respiration, an imbalance in redox homeostasis, reduced GSIS, and reduced β-cell survival in human and rodent β-cells. PAK1 is critical for the formation of the mitochondrial ETC super complexes (CI-CIII and CI-CIII-CIV) and mitochondrial ETC function by binding to and regulating NDUFA12, the CI assembly and stabilizing subunit. Given that PAK1 protein plays two distinct roles, as a kinase, and as a protein scaffold, further investigations will require evaluation of these roles independently, using kinase dead mutants and kinase inactivating compounds, such as IPA3. Further understanding the PAK1 select signaling and scaffolding requirements will inform the downstream pathways that promote mitochondrial function and β-cell survival [123].

5.2. Abnormal Islet β-Cell Mitochondrial Morphology and Dynamics

Abnormal mitochondrial morphology is observed in cadaveric T2D donor β-cells [159]. Distorted mitochondria appear morphologically round and swollen, which impacts mitochondrial function, culminating in impaired GSIS [160,161,162]. Beta cells from T2D and non-diabetic human donors possess similar mitochondrial numbers but T2D donors exhibit significantly higher mitochondrial volume density [159,163]). Several defects in the activation of key enzymes that link glycolysis with tricarboxylic acid cycle have been identified. Each mitochondrion exhibits an increased workload in the T2D milieu, leading to high membrane potential, ROS formation and reduced mtDNA copy number(reviewed in [163]. Under T2D milieu conditions, β-cell mitochondria lose their ability to undergo fusion and become fragmented, leading to β-cell apoptosis [164].
In Goto-Kakizaki (GK) rats, a genetic-lean rat model that exhibits a prediabetes phase similar to humans prior to T2D onset, pancreatic islet β-cells showed mitochondrial network disintegration [165]. In addition, Zucker diabetic fatty rats and HFD-induced models of prediabetes showed mitochondrial remodeling, altered cristae structure, and increased β-cell apoptosis [166,167]. In a model of lean T2D, the Cohen diabetic sensitive (CD) rat, a hyperglycemic model, exhibited markedly diminished β-cell mitochondrial COX activity, defective mitochondria, and increased ROS levels without IR [168,169,170].

5.3. Mitophagy and Elevating Islet β-Cell Dysfunction

Emerging studies of dysregulatory metabolic states representative of prediabetes and T2D are identifying modulation of mitophagy as having a beneficial impact of mitigating disease progression in tissues such as pancreatic β-cells. Recently, it was suggested that mitophagy-deficient β-cells were susceptible to inflammatory stress and overexpressing the mitophagy regulator CLEC16A could protect β-cells from stress [171]. Natural antioxidant compounds have been known to modulate mitophagy and restore mitochondrial dynamics in T2D. However, the efficacy of mitophagy induction by these agents is limited and future studies exploring new avenues are needed, given the fact that mitochondrial dynamics are also important for normal β-cell function. Restoring canonical PARKIN-mediated mitophagy was shown to improve β-cell survival [91,172]. Furthermore, beneficial therapeutic effects in β-cells were observed with selenium nanodots, a synthetic prodrug of antioxidant glutathione peroxidase [173], suggesting the utility of restoring mitophagy to ameliorate oxidative stress in T2D models. While it is still under debate as to whether elevating mitophagy is the appropriate strategy to counteract chronic oxidative stress damage, there is consensus regarding the concept of balancing mitophagy and mitochondrial biogenesis to restore β-cell function.

5.4. Mitochondrial-Endoplasmic Reticulum Miscommunication in Islet β-Cell Dysfunction

Recent evidence demonstrated an association between ER-mitochondrial miscommunication and β-cell dysfunction in individuals with T2D (reviewed in [174,175]. Mitochondria and ER are in close proximity in the β-cell, and they interact via protein bridges and are highly dynamic, forming specific microdomains termed mitochondria-associated membranes (MAM) (reviewed in [176]. It is well established that MAM play a key role in cellular Ca2+ homeostasis [177]. MAM are implicated in regulating mitochondrial dynamics and function, oxidative metabolism, and apoptosis (reviewed in [178]). Furthermore, induction of ER stress by toxic lipids such as palmitate significantly reduced ER-mitochondria crosstalk and altered GSIS in the murine MIN6-B1 β-cell line [175]. Additionally, alterations in ER-mitochondrial crosstalk disrupts lipid homeostasis and is associated with T2D pathogenesis [175]. MAMs can also contain enzymes required for cholesterol and ceramide biosynthesis; there are several reports thus far suggesting elevated ceramide in T2D human islets [166,179]. Ceramide significantly alters the mitochondrial membrane potential (ΔψM), which can trigger the release of proapoptotic factors like cytochrome C, which activates pathways that culminate in β-cell apoptosis [180]. Disruption to MAM assembly compromises mitochondrial dynamics and bioenergetics, and promotes ROS production resulting in the release of mitochondrial-derived proapoptotic factors that promote β-cell death (reviewed in [178]).
We recently revealed that PAK1 enrichment in T2D human islets attenuated markers of ER stress—phosphorylated eukaryotic initiation factor-2α (eIF2α) and C/EBP Homologous Protein (CHOP), and improved β-cell GSIS function [123]. Given also the requirement for PAK1 in β-cell mitochondrial function [123], it remains to be determined if PAK1 functions in the MAM. Taken together, MAM plays a crucial role in dysfunctional β-cells and targeting MAM could be a potential interventional strategy for T2D.

6. Inter-Organ Miscommunications and Oxidative Stress in Insulin Resistance

Conditioned medium experiments and parabiosis experiments have revealed the presence of factors that facilitate inter-tissue/organ communication with other organs/tissues. For example, we recently discovered factor(s) released by skeletal muscle into conditioned medium, that when applied to β-cells, can significantly boost islet β-cell function [121]. Moreover, factors can be shed as soluble factors, such as many hormones (e.g., insulin), and it is now widely appreciated that the vast variety of factors can be shed encased in nanoscale structures.
Cells continuously shed extracellular vesicles (EVs), which are nanoscale bi-lipid structures (~35–1000 nm) that carry biomoleculer cargo, lack a functional nucleus and replicative capacity [181]. EVs have the potential to play a pivotal role as mediators of IR and pancreatic β-cell mass failure in T2D [182,183]; elevated glucose levels, inflammation, increased fatty acids, all of which contribute to oxidative stress, can lead to alterations in the quantity and composition of EVs released from the skeletal muscle, adipose, and pancreas (reviewed in [184]). This section will highlight the emerging role of EVs and a new class of mitochondria-containing EVs (Mito-EVs) [181] in promoting inter-tissue miscommunication during prediabetes and T2D.

6.1. Skeletal Muscle-Derived EVs-Islet β-Cell Crosstalk

Skeletal muscle-derived EVs, possibly containing myokines and microRNAs (miRNAs) [185,186]), have the intriguing ability to transfer lipid-induced IR between muscle cells and β-cells [187,188]. Further, in vivo, EVs from lipid-induced IR muscles can be incorporated into recipient pancreatic β-cells and affect changes to gene expression in those β-cells [187]. Muscle contraction-induced myokines were recently shown to influence glucose uptake, insulin sensitivity and fat metabolism (reviewed in [189]). Furthermore, these secreted myokines from contracted skeletal muscle, including C-X-C motif chemokine ligand 10 (CXCL10), irisin, chemokine ligand 1 (CX3CL1), follistatin and miR-133a (Table 1) also impact the function and survival of β-cells (Figure 6). Thus far, there is potential for skeletal muscle-derived miR-133a to target β-cell mitochondrial UCP2 and influence insulin secretion, given that an increase in skeletal muscle miR-133a leads to reduced UCP2 levels [190] and that β-cell UCP2 controls ROS production to promote GSIS [191]. Indeed, a marked rise in plasma EV protein content, especially antioxidant protein cargo (superoxide dismutase, catalase, and peroxiredoxin) is observed post-exercise to maintain a balanced redox state [192,193]. Currently it is unknown if EV-containing myokines influence pancreatic β-cell mitochondrial function and morphology within the context of prediabetes.

6.2. Adipose-Islet β-Cell Crosstalk and Mitochondrial Transfer

Recently, β-cells and adipose tissue were demonstrated to shed Mito-EVs under pathophysiological conditions (Table 2). There is potential for adipose-derived Mito-EVs to negatively influence β-cells, impairing their functionality in obesity-induced IR/T2D. For example, MIN6 β-cells exposed to EVs derived from mitochondrial stressed adipocytes obtained from caveolin-1 knockout mice, displayed significant cell death and dysfunction [206]. Although the content of these EVs was not examined for mitochondrial constituents, this provides proof of concept implicating inflamed adipocytes in performing detrimental horizontal transfer via EVs. By contrast, adipose-derived Mito-EVs determined to be carrying select mitochondrial proteins (PGM1, PCX and MDH1) positively impacted β-cells: β-cells exposed to Mito-EVs derived from epidymal white adipose tissue (eWAT) of diet-induced obese (DIO) mice displayed increased GSIS [207]. Moreover, in vivo intraperitoneal administration of these Mito-EVs, under fasting conditions, increased plasma insulin content and glucose tolerance in recipient mice, while blockade of EV production with GW4869 prevented glucose tolerance improvements [207]. Interestingly, β-cells exposed to Mito-EVs containing the mitochondrial matrix protein SLC25A5 shed by eWAT adipocytes from lean mice displayed loss of GSIS function [207]. Collectively, these examples point to the need for in-depth characterization of these Mito-EVs and their molecular cargo that carries the capacity to significantly impact β-cell function (Figure 7).
Recent findings also point to β-cells shedding EVs that may have relevance to mitochondrial function. A protein lacking in abundance in T2D tissues, DOC2B (Double C2 domain-containing protein B), was implicated in mitochondrial function and structure in cancer cells [208]; our preliminary findings indicate that β-cells package DOC2B into the lumen of secreted EVs, by virtue of DOC2B’s tandem C2 domains [209]. These roles and locales for DOC2B represent a departure from the classic role for DOC2B as an exocytosis regulatory protein localized to the cytosol. Intriguingly, this finding sparks the question as to whether DOC2B’s requirement for β-cell GSIS stems from a role in the mitochondria, and the potential of DOC2B-laden EVs to serve as indicators of mitochondrial status in β-cells. Advances in EV technologies have deeply improved our understanding of EV mechanisms. However, due to the inherent disadvantages in current EV isolation methods available, distinguishing EVs sources from a specific tissue remains largely a challenge due to the overlap in biophysical properties like size, density and morphology as well as heterogenous protein expression. Future studies may consider targeting β-cell specific proteins to gain insight into the molecular signatures of β-cell-derived EVs observed with mitochondrial dysfunction in T2D.

6.3. Adipose-Immune Cell Crosstalk and Mitochondrial Transfer

Chronic inflammation associated with hyperglycemia and obesity can contribute to increased oxidative stress [210]. Adipose-derived EVs, which includes the subset of Mito-EVs, are implicated in influencing monocyte differentiation into pro-inflammatory macrophages (reviewed in [211]). Adipocyte-to-macrophage mitochondrial transfer serves a regulatory role in systemic metabolic homeostasis [212], and dysregulations in this process may lead to obesity-induced IR/T2D. For example, EVs derived from ob/ob mouse visceral adipocytes (ob-Mito-EVs), whose cargo was partly constituted of mitochondrial proteins HSP70 and HSP90, influence the differentiation of monocytes into macrophages, carrying increased levels of proinflammatory cytokines IL-6 and TNF-α [205]. In vitro, macrophages preincubated with ob-Mito-EVs negatively impact insulin signaling and insulin-stimulated glucose uptake in myocytes compared to control (Figure 7) [205]. In vivo, ob-Mito-EVs led to the development of glucose intolerance and IR in control mice [205]. In contrast, adipose-derived stem-cell (ADSC)-derived EVs (ADSC-EVs) educate macrophages towards an anti-inflammatory polarized M2 macrophage fate by upregulating Arg-1 and IL-10, without significantly increasing pro-inflammatory M1-related iNOS, TNF-α, and IL-12 [213]. Indeed, HFD-obese mice treated with ADSC-EVs showed significantly improved glucose tolerance and insulin sensitivity [213]. One caveat though is that the cargo in these ADSC-EVs was not characterized for mitochondrial proteins per se. Nevertheless, these data suggest that the cargo within the adipocyte EVs influences monocyte/macrophage signaling in insulin-sensitive tissues (Figure 7).

7. Antioxidant Interventions for Prediabetes and T2D

Currently, a healthy diet consisting of abundant plant-based foods is a primary intervention for prediabetes individuals. Given the association between oxidative stress and T2D, a plethora of preclinical and clinical investigations into the therapeutic potential of natural antioxidants (i.e., polyphenols and vitamins) are currently underway to mitigate disease progression. Below we will focus on the leading natural antioxidant molecules that are substantially researched in clinical trials for T2D.
Curcumin: Curcumin is a natural polyphenol that is derived from the turmeric plant Curcuma longa. In vitro, curcumin inhibits ROS production and NF-κB activation in IR C2C12 myotubes, demonstrating antioxidant and anti-inflammatory effects with potential benefits for treating T2D [214]. However, clinical trials and meta-analyses have reported mixed results. One meta-analysis capturing 72% of clinical trials showed curcumin improving fasting blood glucose levels and hemoglobin A1c (HbA1c), whilst only three out of the five studies showed a decrease in Homeostatic Model Assessment for Insulin Resistance (HOMA-IR) [215]. However, curcumin treatment had beneficial impact on improving antioxidant capacity and alleviating oxidative stress. Furthermore, a chemically-modified curcuminoid showed improved insulin tolerance, fasting glycemia and alleviated oxidative stress in GK rats [216]. Thus, future studies utilizing synthetic derivatives of curcumin could maximize the therapeutic efficacy in T2D.
Resveratrol: Resveratrol is a polyphenol with free radical- and H2O2-scavenging capabilities that is naturally found in plant-based foods. IR rat myotubes studies indicated that resveratrol could restore glucose uptake and fatty acid oxidation via AMPK mediated pathways [217,218,219]. In addition, T2D GK non-obese rats receiving 20 mg/kg/day of resveratrol [220], and HFD-fed C57BL6/J male mice receiving 100 mg/kg/day of resveratrol [219], demonstrated improvements in IR and lipid metabolism and diminishments in muscle inflammation. In a placebo-controlled study, T2D patients receiving 800 mg/day resveratrol for 2 months showed reduced protein carbonyl content and white blood cell O2•− levels and increased total thiol content, showing improved antioxidant capacity [221]. However, meta-analyses revealed mixed results for resveratrol safety and effectiveness in T2D [222,223,224] and thus, adequately powered randomized controlled trials are needed.
Vitamin C: Vitamin C is a naturally occurring antioxidant in citrus-based fruits. The recommended adult dietary intake of vitamin C ranges from 40 to 110 mg/day (reviewed in [225]. Vitamin C supplementation ranging from 500 to 1000 mg/day in T2D patients has demonstrated diminished production of ROS/RNS, improved antioxidant capacity, and improved glycemic control [225,226,227,228]. However, the high heterogeneity for glycemic control and small sample sizes [225] limit the conclusions that can be drawn regarding vitamin C treatment for T2D.
Vitamin D: Vitamin D is a lipid-soluble secosteroid; vitamin D3 is the animal-derived version formed from 7-dehydrocholsterol in the skin from natural sunlight. Vitamin D functions in genomic regulation via its receptor, and regulates signaling pathways that include MAP kinase and protein kinase C. In skeletal muscle, vitamin D treatment is associated with improved insulin sensitivity, reversal of myosteatosis, and decreased inflammation activation via NF-κB in HFD-fed mice [229]. Oxidative stress is inhibited in vitamin D-treated C2C12 muscle cells [230], suggesting a therapeutic role for vitamin D in T2D. A meta-analysis revealed that T2D patients receiving vitamin D supplementation to correct a vitamin D deficiency had reductions in HbA1c levels, fasting glucose and HOMA-IR [231]. Furthermore, placebo-controlled trials showed that vitamin D supplementation (4000 IU per day) lowers the risk of T2D progression in prediabetic individuals [232]. High doses of vitamin D at 100,000 or 200,000 IU biweekly over 6 months was associated with a reduction in oxidative stress [233]. Vitamin D supplementation may also be useful at decreasing inflammatory markers MCP-1 and IL-8 and attenuating oxidative stress in T2D [234]. Intriguingly, other natural antioxidant therapeutic candidates (e.g., zinc, omega-3 fatty acids, and vitamin E) have not shown any impact on glycemic control suggesting that vitamin D may be a leading antioxidant candidate for T2D [235].

8. Perspectives: Mitochondria and ROS as Therapeutic Targets for Prediabetes

One long-standing debate within the metabolic research field is whether mitochondrial dysfunction contributes to skeletal muscle IR, or is merely a bystander. A recent study that advocated the bystander explanation reported no difference in mitochondrial dysfunction between skeletal muscle biopsies from non-diabetic obese and T2D obese cohorts [236]. However, the non-diabetic obese cohort [236] had HbA1c levels that overlapped with the established prediabetic range of 5.7–6.4% [115]. In contrast, a recent proteomics study on skeletal muscle biopsies from T2D, prediabetic or non-obese human males identified decreases in mitochondrial proteins during prediabetes, which worsened upon progression to T2D [115]. These findings highlight the need to distinguish the mitochondrial activity patterns between T2D and prediabetes.
It is perhaps inadequate to solely focus on mitochondrial activity when assessing mitochondrial dysfunction. Evidence for subcellular mitochondrial heterogeneity (i.e., variance in mitochondrial ultrastructure and dynamics) is emerging in islet β-cells [237], and metabolic dysregulation can increase this heterogeneity [164]. Analyzing the degree to which mitochondrial heterogeneity contributes to skeletal muscle IR and β-cell dysfunction within the four T2D archetypes should further contribute to developing different therapeutic strategies for targeting mitochondrial dysfunction in prediabetes.
No drug treatments in the United States are currently approved for prediabetes, although off-label prescribing of metformin is recommended for select individuals with prediabetes [238]. Of relevance to the focus of this review, emerging evidences suggest that metformin reduces mitochondrial dysfunction in T2D individuals by reducing mROS production, preventing pro-fission mitochondrial morphology, boosting the antioxidant response, and inducing an increase in mitophagy [239,240]. Consistent with this, metformin reduces the mitochondrial pro-fission phenotype via AMPK in diabetic mice [241]. It will be beneficial to determine the prevalence of metformin use and the degree to which who is benefiting from metformin use in the prediabetes clinic setting given the heterogenous treatment effects reported for T2D [242]. Furthermore, there are serious side-effects associated with metformin use (e.g., lactic acidosis) and tolerability is individualistic. Thus, new safe and effective therapies that target novel mitochondrial dysfunction pathways and has a high benefit: risk ratio for prediabetes are needed.
Focusing on the prediabetes stage may also provide insights into the potential for effectiveness of antioxidant supplements at halting disease progression. T2D disease heterogeneity and the degree of oxidative stress are currently limiting factors in assessing antioxidant supplement efficacy to mitigate an individual’s oxidative stress. For example, three analyses indicate differential conclusions: (1) the Bialystock PLUS population identified a trend of lower total antioxidant capacity in individuals with prediabetes as compared to healthy controls in middle-aged groups [30]; (2) a meta-analysis encompassing 11 studies across Asia with different polyphenol, vitamin and mineral supplementation showed improvements with IR was correlated with total antioxidant capacity improvements [31]; (3) the 2019 Rotterdam study that evaluated antioxidant capacity using the nutritional gold standard of ferric reducing ability of plasma (FRAP) assay [28] across 5796 individuals identified that even though multi-antioxidant dietary rich intake lowered the risks of T2D per HOMA-IR scores, the risk of prediabetes was not reduced [29]. Furthermore, suboptimal dietary intakes (i.e., low antioxidant rich foods) were shown to be responsible for 7/10 newly diagnosed T2D cases across and within the populations of 184 high- and low-income countries [27]. Thus, the natural heterogeneity of access to antioxidant rich diets within various populations and geographic locations, as well as individual differences need to be taken into account when evaluating antioxidant therapeutic efficacy.
In summary, significant progress has been made thus far in the metabolic field. The body of evidence supports the notion that (1) T2D is heterogenous in nature and IR does not necessarily precede T2D development, (2) oxidative damage occurs early during disease onset, (3) boosting mitophagy could improve islet β-cell and skeletal muscle function, and (4) cargo shed from metabolically stressed tissues may contribute towards T2D progression. Further research is needed to understand and elucidate the regulation and relevance of ROS within the various subcellular microdomains in health and metabolic dysfunction. Measurement of ROS in the subcellular microdomain may help overcome some of the challenges faced when developing new therapies dealing with the heterogeneity that exists within T2D. Currently, individuals are stratified based on BMI for therapeutic interventions, yet due to phenotypic heterogeneity, this does not capture all in this “one-size-fits-all” approach. Future studies may consider targeting ROS and mitochondrial dysfunction, as a means to treat more T2D archetypes versus obese-T2D or lean-T2D. Future focus on developing synthetic derivatives of natural antioxidants, multi-faceted therapies that dually promote pancreatic and skeletal muscle glucose metabolism and mitochondrial function, expanding clinical studies to include a focus on prediabetes and taking individual antioxidant capacities into account may provide the best strategy for preventing T2D progression.

Author Contributions

D.C.T. and R.V. Conceptualization, Writing—review and editing. D.E., J.M.H., R.B., M.A., E.O. and C.S.J. Writing—original draft and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by grants from the National Institutes of Health (DK067912 and DK112917 to D.C.T., and DK102233 to D.C.T. and R.V.), the Juvenile Diabetes Research Foundation (17-2013-454 to D.C.T.), fellowships from the American Heart Association (#905770, to R.B.), and the Ford Foundation (predoctoral, to D.E.), and the Wanek Project to Cure Type 1 Diabetes at the City of Hope to D.C.T.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cefalu, W.T.; Andersen, D.K.; Arreaza-Rubín, G.; Pin, C.L.; Sato, S.; Verchere, C.B.; Woo, M.; Rosenblum, N.D. Heterogeneity of Diabetes: β-Cells, Phenotypes, and Precision Medicine: Proceedings of an International Symposium of the Canadian Institutes of Health Research’s Institute of Nutrition, Metabolism and Diabetes and the U.S. National Institutes of Health’s National Institute of Diabetes and Digestive and Kidney Diseases. Diabetes Care 2022, 45, 3–22. [Google Scholar] [CrossRef] [PubMed]
  2. International Diabetes Federation. Diabetes Atlas, 10th ed.; International Diabetes Federation: Brussels, Belgium, 2021. [Google Scholar]
  3. American Diabetes Association Professional Practice Committee. 2. Diagnosis and Classification of Diabetes: Standards of Care in Diabetes—2024. Diabetes Care 2023, 47, S20–S42. [Google Scholar] [CrossRef]
  4. Wesolowska-Andersen, A.; Brorsson, C.A.; Bizzotto, R.; Mari, A.; Tura, A.; Koivula, R.; Mahajan, A.; Vinuela, A.; Tajes, J.F.; Sharma, S.; et al. Four groups of type 2 diabetes contribute to the etiological and clinical heterogeneity in newly diagnosed individuals: An IMI DIRECT study. Cell Rep. Med. 2022, 3, 100477. [Google Scholar] [CrossRef] [PubMed]
  5. Kodama, K.; Tojjar, D.; Yamada, S.; Toda, K.; Patel, C.J.; Butte, A.J. Ethnic differences in the relationship between insulin sensitivity and insulin response: A systematic review and meta-analysis. Diabetes Care 2013, 36, 1789–1796. [Google Scholar] [CrossRef] [PubMed]
  6. Vicks, W.S.; Lo, J.C.; Guo, L.; Rana, J.S.; Zhang, S.; Ramalingam, N.D.; Gordon, N.P. Prevalence of prediabetes and diabetes vary by ethnicity among U.S. Asian adults at healthy weight, overweight, and obesity ranges: An electronic health record study. BMC Public Health 2022, 22, 1954. [Google Scholar] [CrossRef] [PubMed]
  7. Staimez, L.R.; Weber, M.B.; Ranjani, H.; Ali, M.K.; Echouffo-Tcheugui, J.B.; Phillips, L.S.; Mohan, V.; Narayan, K.M. Evidence of reduced beta-cell function in Asian Indians with mild dysglycemia. Diabetes Care 2013, 36, 2772–2778. [Google Scholar] [CrossRef] [PubMed]
  8. Ke, C.; Narayan, K.M.V.; Chan, J.C.N.; Jha, P.; Shah, B.R. Pathophysiology, phenotypes and management of type 2 diabetes mellitus in Indian and Chinese populations. Nat. Rev. Endocrinol. 2022, 18, 413–432. [Google Scholar] [CrossRef]
  9. Kanaya, A.M.; Herrington, D.; Vittinghoff, E.; Ewing, S.K.; Liu, K.; Blaha, M.J.; Dave, S.S.; Qureshi, F.; Kandula, N.R. Understanding the high prevalence of diabetes in U.S. south Asians compared with four racial/ethnic groups: The MASALA and MESA studies. Diabetes Care 2014, 37, 1621–1628. [Google Scholar] [CrossRef]
  10. Gopaul, N.K.; Manraj, M.D.; Hebe, A.; Lee Kwai Yan, S.; Johnston, A.; Carrier, M.J.; Anggard, E.E. Oxidative stress could precede endothelial dysfunction and insulin resistance in Indian Mauritians with impaired glucose metabolism. Diabetologia 2001, 44, 706–712. [Google Scholar] [CrossRef]
  11. Urakawa, H.; Katsuki, A.; Sumida, Y.; Gabazza, E.C.; Murashima, S.; Morioka, K.; Maruyama, N.; Kitagawa, N.; Tanaka, T.; Hori, Y.; et al. Oxidative stress is associated with adiposity and insulin resistance in men. J. Clin. Endocrinol. Metab. 2003, 88, 4673–4676. [Google Scholar] [CrossRef]
  12. Houstis, N.; Rosen, E.D.; Lander, E.S. Reactive oxygen species have a causal role in multiple forms of insulin resistance. Nature 2006, 440, 944–948. [Google Scholar] [CrossRef] [PubMed]
  13. Hudish, L.I.; Reusch, J.E.; Sussel, L. Beta Cell dysfunction during progression of metabolic syndrome to type 2 diabetes. J. Clin. Investig. 2019, 129, 4001–4008. [Google Scholar] [CrossRef]
  14. Sutton, E.F.; Beyl, R.; Early, K.S.; Cefalu, W.T.; Ravussin, E.; Peterson, C.M. Early Time-Restricted Feeding Improves Insulin Sensitivity, Blood Pressure, and Oxidative Stress Even without Weight Loss in Men with Prediabetes. Cell Metab. 2018, 27, 1212–1221.e3. [Google Scholar] [CrossRef]
  15. Dimova, R.; Chakarova, N.; Grozeva, G.; Kirilov, G.; Tankova, T. The relationship between glucose variability and insulin sensitivity and oxidative stress in subjects with prediabetes. Diabetes Res. Clin. Pract. 2019, 158, 107911. [Google Scholar] [CrossRef] [PubMed]
  16. Samocha-Bonet, D.; Heilbronn, L.K.; Lichtenberg, D.; Campbell, L.V. Does skeletal muscle oxidative stress initiate insulin resistance in genetically predisposed individuals? Trends Endocrinol. Metab. 2010, 21, 83–88. [Google Scholar] [CrossRef] [PubMed]
  17. Monserrat-Mesquida, M.; Quetglas-Llabres, M.; Capo, X.; Bouzas, C.; Mateos, D.; Pons, A.; Tur, J.A.; Sureda, A. Metabolic Syndrome is Associated with Oxidative Stress and Proinflammatory State. Antioxidants 2020, 9, 236. [Google Scholar] [CrossRef] [PubMed]
  18. Murphy, M.P.; Bayir, H.; Belousov, V.; Chang, C.J.; Davies, K.J.A.; Davies, M.J.; Dick, T.P.; Finkel, T.; Forman, H.J.; Janssen-Heininger, Y.; et al. Guidelines for measuring reactive oxygen species and oxidative damage in cells and in vivo. Nat. Metab. 2022, 4, 651–662. [Google Scholar] [CrossRef] [PubMed]
  19. Abdelbaset-Ismail, A.; Ciechanowicz, A.K.; Bujko, K.; Ratajczak, J.; Kucia, M.; Ratajczak, M.Z. The Nox2-ROS-Nlrp3 Inflammasome Signaling Stimulates in the Hematopoietic Stem/Progenitor Cells Lipogenesis to Facilitate Membrane Lipid Raft Formation. Stem Cell Rev. Rep. 2023, 19, 92–103. [Google Scholar] [CrossRef]
  20. Li, Y.; Liu, C.; Rolling, L.; Sikora, V.; Chen, Z.; Gurwin, J.; Barabell, C.; Lin, J.; Duan, C. ROS signaling-induced mitochondrial Sgk1 expression regulates epithelial cell renewal. Proc. Natl. Acad. Sci. USA 2023, 120, e2216310120. [Google Scholar] [CrossRef]
  21. Pi, J.; Bai, Y.; Zhang, Q.; Wong, V.; Floering, L.M.; Daniel, K.; Reece, J.M.; Deeney, J.T.; Andersen, M.E.; Corkey, B.E.; et al. Reactive oxygen species as a signal in glucose-stimulated insulin secretion. Diabetes 2007, 56, 1783–1791. [Google Scholar] [CrossRef]
  22. Plecita-Hlavata, L.; Jaburek, M.; Holendova, B.; Tauber, J.; Pavluch, V.; Berkova, Z.; Cahova, M.; Schroder, K.; Brandes, R.P.; Siemen, D.; et al. Glucose-Stimulated Insulin Secretion Fundamentally Requires H2O2 Signaling by NADPH Oxidase 4. Diabetes 2020, 69, 1341–1354. [Google Scholar] [CrossRef]
  23. Merry, T.L.; Steinberg, G.R.; Lynch, G.S.; McConell, G.K. Skeletal muscle glucose uptake during contraction is regulated by nitric oxide and ROS independently of AMPK. Am. J. Physiol. Endocrinol. Metab. 2010, 298, E577–E585. [Google Scholar] [CrossRef]
  24. Henriquez-Olguin, C.; Knudsen, J.R.; Raun, S.H.; Li, Z.; Dalbram, E.; Treebak, J.T.; Sylow, L.; Holmdahl, R.; Richter, E.A.; Jaimovich, E.; et al. Cytosolic ROS production by NADPH oxidase 2 regulates muscle glucose uptake during exercise. Nat. Commun. 2019, 10, 4623. [Google Scholar] [CrossRef]
  25. Xirouchaki, C.E.; Jia, Y.; McGrath, M.J.; Greatorex, S.; Tran, M.; Merry, T.L.; Hong, D.; Eramo, M.J.; Broome, S.C.; Woodhead, J.S.T.; et al. Skeletal muscle NOX4 is required for adaptive responses that prevent insulin resistance. Sci. Adv. 2021, 7, eabl4988. [Google Scholar] [CrossRef]
  26. Dikalov, S.I.; Nazarewicz, R.R.; Bikineyeva, A.; Hilenski, L.; Lassegue, B.; Griendling, K.K.; Harrison, D.G.; Dikalova, A.E. Nox2-induced production of mitochondrial superoxide in angiotensin II-mediated endothelial oxidative stress and hypertension. Antioxid. Redox Signal. 2014, 20, 281–294. [Google Scholar] [CrossRef]
  27. Kroller-Schon, S.; Steven, S.; Kossmann, S.; Scholz, A.; Daub, S.; Oelze, M.; Xia, N.; Hausding, M.; Mikhed, Y.; Zinssius, E.; et al. Molecular mechanisms of the crosstalk between mitochondria and NADPH oxidase through reactive oxygen species-studies in white blood cells and in animal models. Antioxid. Redox Signal. 2014, 20, 247–266. [Google Scholar] [CrossRef]
  28. Ashcroft, F.M.; Harrison, D.E.; Ashcroft, S.J. Glucose induces closure of single potassium channels in isolated rat pancreatic beta-cells. Nature 1984, 312, 446–448. [Google Scholar] [CrossRef] [PubMed]
  29. Cook, D.L.; Hales, C.N. Intracellular ATP directly blocks K+ channels in pancreatic B-cells. Nature 1984, 311, 271–273. [Google Scholar] [CrossRef] [PubMed]
  30. Campbell, J.E.; Newgard, C.B. Mechanisms controlling pancreatic islet cell function in insulin secretion. Nat. Rev. Mol. Cell Biol. 2021, 22, 142–158. [Google Scholar] [CrossRef] [PubMed]
  31. Zhao, R.Z.; Jiang, S.; Zhang, L.; Yu, Z.B. Mitochondrial electron transport chain, ROS generation and uncoupling (Review). Int. J. Mol. Med. 2019, 44, 3–15. [Google Scholar] [CrossRef]
  32. Camargo, L.L.; Harvey, A.P.; Rios, F.J.; Tsiropoulou, S.; Da Silva, R.N.O.; Cao, Z.; Graham, D.; McMaster, C.; Burchmore, R.J.; Hartley, R.C.; et al. Vascular Nox (NADPH Oxidase) Compartmentalization, Protein Hyperoxidation, and Endoplasmic Reticulum Stress Response in Hypertension. Hypertension 2018, 72, 235–246. [Google Scholar] [CrossRef] [PubMed]
  33. Block, K.; Gorin, Y.; Abboud, H.E. Subcellular localization of Nox4 and regulation in diabetes. Proc. Natl. Acad. Sci. USA 2009, 106, 14385–14390. [Google Scholar] [CrossRef] [PubMed]
  34. Shanmugasundaram, K.; Nayak, B.K.; Friedrichs, W.E.; Kaushik, D.; Rodriguez, R.; Block, K. NOX4 functions as a mitochondrial energetic sensor coupling cancer metabolic reprogramming to drug resistance. Nat. Commun. 2017, 8, 997. [Google Scholar] [CrossRef] [PubMed]
  35. Rodriguez-Nuevo, A.; Torres-Sanchez, A.; Duran, J.M.; De Guirior, C.; Martinez-Zamora, M.A.; Boke, E. Oocytes maintain ROS-free mitochondrial metabolism by suppressing complex I. Nature 2022, 607, 756–761. [Google Scholar] [CrossRef] [PubMed]
  36. Datta, K.; Suman, S.; Kallakury, B.V.; Fornace, A.J., Jr. Exposure to heavy ion radiation induces persistent oxidative stress in mouse intestine. PLoS ONE 2012, 7, e42224. [Google Scholar] [CrossRef] [PubMed]
  37. Pascoe, C.D.; Roy, N.; Turner-Brannen, E.; Schultz, A.; Vaghasiya, J.; Ravandi, A.; Halayko, A.J.; West, A.R. Oxidized phosphatidylcholines induce multiple functional defects in airway epithelial cells. Am. J. Physiol. Lung Cell Mol. Physiol. 2021, 321, L703–L717. [Google Scholar] [CrossRef] [PubMed]
  38. Li, N.; Brun, T.; Cnop, M.; Cunha, D.A.; Eizirik, D.L.; Maechler, P. Transient oxidative stress damages mitochondrial machinery inducing persistent beta-cell dysfunction. J. Biol. Chem. 2009, 284, 23602–23612. [Google Scholar] [CrossRef]
  39. Urbano, F.; Bugliani, M.; Filippello, A.; Scamporrino, A.; Di Mauro, S.; Di Pino, A.; Scicali, R.; Noto, D.; Rabuazzo, A.M.; Averna, M.; et al. Atorvastatin but Not Pravastatin Impairs Mitochondrial Function in Human Pancreatic Islets and Rat beta-Cells. Direct Effect of Oxidative Stress. Sci. Rep. 2017, 7, 11863. [Google Scholar] [CrossRef]
  40. Vanhorebeek, I.; De Vos, R.; Mesotten, D.; Wouters, P.J.; De Wolf-Peeters, C.; Van den Berghe, G. Protection of hepatocyte mitochondrial ultrastructure and function by strict blood glucose control with insulin in critically ill patients. Lancet 2005, 365, 53–59. [Google Scholar] [CrossRef]
  41. Kelley, D.E.; He, J.; Menshikova, E.V.; Ritov, V.B. Dysfunction of mitochondria in human skeletal muscle in type 2 diabetes. Diabetes 2002, 51, 2944–2950. [Google Scholar] [CrossRef]
  42. Vincent, A.M.; Brownlee, M.; Russell, J.W. Oxidative stress and programmed cell death in diabetic neuropathy. Ann. N. Y. Acad. Sci. 2002, 959, 368–383. [Google Scholar] [CrossRef] [PubMed]
  43. Scott, L.J.; Erdos, M.R.; Huyghe, J.R.; Welch, R.P.; Beck, A.T.; Wolford, B.N.; Chines, P.S.; Didion, J.P.; Narisu, N.; Stringham, H.M.; et al. The genetic regulatory signature of type 2 diabetes in human skeletal muscle. Nat. Commun. 2016, 7, 11764. [Google Scholar] [CrossRef] [PubMed]
  44. Mootha, V.K.; Lindgren, C.M.; Eriksson, K.F.; Subramanian, A.; Sihag, S.; Lehar, J.; Puigserver, P.; Carlsson, E.; Ridderstråle, M.; Laurila, E.; et al. PGC-1alpha-responsive genes involved in oxidative phosphorylation are coordinately downregulated in human diabetes. Nat. Genet. 2003, 34, 267–273. [Google Scholar] [CrossRef]
  45. Patti, M.E.; Butte, A.J.; Crunkhorn, S.; Cusi, K.; Berria, R.; Kashyap, S.; Miyazaki, Y.; Kohane, I.; Costello, M.; Saccone, R.; et al. Coordinated reduction of genes of oxidative metabolism in humans with insulin resistance and diabetes: Potential role of PGC1 and NRF1. Proc. Natl. Acad. Sci. USA 2003, 100, 8466–8471. [Google Scholar] [CrossRef] [PubMed]
  46. La Sala, L.; Mrakic-Sposta, S.; Tagliabue, E.; Prattichizzo, F.; Micheloni, S.; Sangalli, E.; Specchia, C.; Uccellatore, A.C.; Lupini, S.; Spinetti, G.; et al. Circulating microRNA-21 is an early predictor of ROS-mediated damage in subjects with high risk of developing diabetes and in drug-naive T2D. Cardiovasc. Diabetol. 2019, 18, 18. [Google Scholar] [CrossRef] [PubMed]
  47. Neal, A.; Rountree, A.; Kernan, K.; Van Yserloo, B.; Zhang, H.; Reed, B.J.; Osborne, W.; Wang, W.; Sweet, I.R. Real-time imaging of intracellular hydrogen peroxide in pancreatic islets. Biochem. J. 2016, 473, 4443–4456. [Google Scholar] [CrossRef] [PubMed]
  48. Llanos, P.; Contreras-Ferrat, A.; Barrientos, G.; Valencia, M.; Mears, D.; Hidalgo, C. Glucose-Dependent Insulin Secretion in Pancreatic beta-Cell Islets from Male Rats Requires Ca2+ Release via ROS-Stimulated Ryanodine Receptors. PLoS ONE 2015, 10, e0129238. [Google Scholar] [CrossRef]
  49. Syed, I.; Kyathanahalli, C.N.; Jayaram, B.; Govind, S.; Rhodes, C.J.; Kowluru, R.A.; Kowluru, A. Increased phagocyte-like NADPH oxidase and ROS generation in type 2 diabetic ZDF rat and human islets: Role of Rac1-JNK1/2 signaling pathway in mitochondrial dysregulation in the diabetic islet. Diabetes 2011, 60, 2843–2852. [Google Scholar] [CrossRef]
  50. Jayaram, B.; Kowluru, A. Phagocytic NADPH oxidase links ARNO-Arf6 signaling pathway in glucose-stimulated insulin secretion from the pancreatic beta-cell. Cell Physiol. Biochem. 2012, 30, 1351–1362. [Google Scholar] [CrossRef]
  51. Leloup, C.; Tourrel-Cuzin, C.; Magnan, C.; Karaca, M.; Castel, J.; Carneiro, L.; Colombani, A.L.; Ktorza, A.; Casteilla, L.; Penicaud, L. Mitochondrial reactive oxygen species are obligatory signals for glucose-induced insulin secretion. Diabetes 2009, 58, 673–681. [Google Scholar] [CrossRef]
  52. Affourtit, C.; Brand, M.D. Uncoupling protein-2 contributes significantly to high mitochondrial proton leak in INS-1E insulinoma cells and attenuates glucose-stimulated insulin secretion. Biochem. J. 2008, 409, 199–204. [Google Scholar] [CrossRef]
  53. Affourtit, C.; Jastroch, M.; Brand, M.D. Uncoupling protein-2 attenuates glucose-stimulated insulin secretion in INS-1E insulinoma cells by lowering mitochondrial reactive oxygen species. Free Radic. Biol. Med. 2011, 50, 609–616. [Google Scholar] [CrossRef]
  54. Martyn, K.D.; Frederick, L.M.; von Loehneysen, K.; Dinauer, M.C.; Knaus, U.G. Functional analysis of Nox4 reveals unique characteristics compared to other NADPH oxidases. Cell. Signal. 2006, 18, 69–82. [Google Scholar] [CrossRef] [PubMed]
  55. Elksnis, A.; Cen, J.; Wikström, P.; Carlsson, P.O.; Welsh, N. Pharmacological Inhibition of NOX4 Improves Mitochondrial Function and Survival in Human Beta-Cells. Biomedicines 2021, 9, 1865. [Google Scholar] [CrossRef] [PubMed]
  56. Stancill, J.S.; Hansen, P.A.; Mathison, A.J.; Schmidt, E.E.; Corbett, J.A. Deletion of Thioredoxin Reductase Disrupts Redox Homeostasis and Impairs β-Cell Function. Function 2022, 3, zqac034. [Google Scholar] [CrossRef] [PubMed]
  57. Stancill, J.S.; Broniowska, K.A.; Oleson, B.J.; Naatz, A.; Corbett, J.A. Pancreatic beta-cells detoxify H2O2 through the peroxiredoxin/thioredoxin antioxidant system. J. Biol. Chem. 2019, 294, 4843–4853. [Google Scholar] [CrossRef] [PubMed]
  58. Baumel-Alterzon, S.; Katz, L.S.; Brill, G.; Jean-Pierre, C.; Li, Y.; Tse, I.; Biswal, S.; Garcia-Ocana, A.; Scott, D.K. Nrf2 Regulates beta-Cell Mass by Suppressing beta-Cell Death and Promoting beta-Cell Proliferation. Diabetes 2022, 71, 989–1011. [Google Scholar] [CrossRef] [PubMed]
  59. DeFronzo, R.A.; Ferrannini, E.; Sato, Y.; Felig, P.; Wahren, J. Synergistic interaction between exercise and insulin on peripheral glucose uptake. J. Clin. Investig. 1981, 68, 1468–1474. [Google Scholar] [CrossRef]
  60. Baron, A.D.; Brechtel, G.; Wallace, P.; Edelman, S.V. Rates and tissue sites of non-insulin- and insulin-mediated glucose uptake in humans. Am. J. Physiol. 1988, 255, E769–E774. [Google Scholar] [CrossRef]
  61. Powers, S.K.; Ji, L.L.; Kavazis, A.N.; Jackson, M.J. Reactive oxygen species: Impact on skeletal muscle. Compr. Physiol. 2011, 1, 941–969. [Google Scholar] [CrossRef]
  62. Higaki, Y.; Mikami, T.; Fujii, N.; Hirshman, M.F.; Koyama, K.; Seino, T.; Tanaka, K.; Goodyear, L.J. Oxidative stress stimulates skeletal muscle glucose uptake through a phosphatidylinositol 3-kinase-dependent pathway. Am. J. Physiol. Endocrinol. Metab. 2008, 294, E889–E897. [Google Scholar] [CrossRef] [PubMed]
  63. Yasuda, T.; Ishihara, T.; Ichimura, A.; Ishihara, N. Mitochondrial dynamics define muscle fiber type by modulating cellular metabolic pathways. Cell Rep. 2023, 42, 112434. [Google Scholar] [CrossRef]
  64. Jani, S.; Da Eira, D.; Hadday, I.; Bikopoulos, G.; Mohasses, A.; de Pinho, R.A.; Ceddia, R.B. Distinct mechanisms involving diacylglycerol, ceramides, and inflammation underlie insulin resistance in oxidative and glycolytic muscles from high fat-fed rats. Sci. Rep. 2021, 11, 19160. [Google Scholar] [CrossRef] [PubMed]
  65. Makinen, S.; Nguyen, Y.H.; Skrobuk, P.; Koistinen, H.A. Palmitate and oleate exert differential effects on insulin signalling and glucose uptake in human skeletal muscle cells. Endocr. Connect. 2017, 6, 331–339. [Google Scholar] [CrossRef] [PubMed]
  66. Barazzoni, R.; Zanetti, M.; Gortan Cappellari, G.; Semolic, A.; Boschelle, M.; Codarin, E.; Pirulli, A.; Cattin, L.; Guarnieri, G. Fatty acids acutely enhance insulin-induced oxidative stress and cause insulin resistance by increasing mitochondrial reactive oxygen species (ROS) generation and nuclear factor-kappaB inhibitor (IkappaB)-nuclear factor-kappaB (NFkappaB) activation in rat muscle, in the absence of mitochondrial dysfunction. Diabetologia 2012, 55, 773–782. [Google Scholar] [CrossRef] [PubMed]
  67. Chan, D.C. Mitochondrial Dynamics and Its Involvement in Disease. Annu. Rev. Pathol. Mech. Dis. 2020, 15, 235–259. [Google Scholar] [CrossRef] [PubMed]
  68. Yu, T.; Sheu, S.S.; Robotham, J.L.; Yoon, Y. Mitochondrial fission mediates high glucose-induced cell death through elevated production of reactive oxygen species. Cardiovasc. Res. 2008, 79, 341–351. [Google Scholar] [CrossRef]
  69. Frank, S.; Gaume, B.; Bergmann-Leitner, E.S.; Leitner, W.W.; Robert, E.G.; Catez, F.; Smith, C.L.; Youle, R.J. The role of dynamin-related protein 1, a mediator of mitochondrial fission, in apoptosis. Dev. Cell 2001, 1, 515–525. [Google Scholar] [CrossRef]
  70. Gundersen, A.E.; Kugler, B.A.; McDonald, P.M.; Veraksa, A.; Houmard, J.A.; Zou, K. Altered mitochondrial network morphology and regulatory proteins in mitochondrial quality control in myotubes from severely obese humans with or without type 2 diabetes. Appl. Physiol. Nutr. Metab. 2020, 45, 283–293. [Google Scholar] [CrossRef]
  71. Espinosa, A.; Campos, C.; Diaz-Vegas, A.; Galgani, J.E.; Juretic, N.; Osorio-Fuentealba, C.; Bucarey, J.L.; Tapia, G.; Valenzuela, R.; Contreras-Ferrat, A.; et al. Insulin-dependent H2O2 production is higher in muscle fibers of mice fed with a high-fat diet. Int. J. Mol. Sci. 2013, 14, 15740–15754. [Google Scholar] [CrossRef]
  72. Dos Santos, J.M.; de Oliveira, D.S.; Moreli, M.L.; Benite-Ribeiro, S.A. The role of mitochondrial DNA damage at skeletal muscle oxidative stress on the development of type 2 diabetes. Mol. Cell. Biochem. 2018, 449, 251–255. [Google Scholar] [CrossRef]
  73. Fink, L.N.; Costford, S.R.; Lee, Y.S.; Jensen, T.E.; Bilan, P.J.; Oberbach, A.; Bluher, M.; Olefsky, J.M.; Sams, A.; Klip, A. Pro-inflammatory macrophages increase in skeletal muscle of high fat-fed mice and correlate with metabolic risk markers in humans. Obesity 2014, 22, 747–757. [Google Scholar] [CrossRef]
  74. Shabani, M.; Sadeghi, A.; Hosseini, H.; Teimouri, M.; Babaei Khorzoughi, R.; Pasalar, P.; Meshkani, R. Resveratrol alleviates obesity-induced skeletal muscle inflammation via decreasing M1 macrophage polarization and increasing the regulatory T cell population. Sci. Rep. 2020, 10, 3791. [Google Scholar] [CrossRef] [PubMed]
  75. Kleinert, M.; Parker, B.L.; Jensen, T.E.; Raun, S.H.; Pham, P.; Han, X.; James, D.E.; Richter, E.A.; Sylow, L. Quantitative proteomic characterization of cellular pathways associated with altered insulin sensitivity in skeletal muscle following high-fat diet feeding and exercise training. Sci. Rep. 2018, 8, 10723. [Google Scholar] [CrossRef] [PubMed]
  76. Merz, K.E.; Hwang, J.; Zhou, C.; Veluthakal, R.; McCown, E.M.; Hamilton, A.; Oh, E.; Dai, W.; Fueger, P.T.; Jiang, L.; et al. Enrichment of the exocytosis protein STX4 in skeletal muscle remediates peripheral insulin resistance and alters mitochondrial dynamics via Drp1. Nat. Commun. 2022, 13, 424. [Google Scholar] [CrossRef] [PubMed]
  77. Sergi, D.; Luscombe-Marsh, N.; Naumovski, N.; Abeywardena, M.; O’Callaghan, N. Palmitic Acid, but Not Lauric Acid, Induces Metabolic Inflammation, Mitochondrial Fragmentation, and a Drop in Mitochondrial Membrane Potential in Human Primary Myotubes. Front. Nutr. 2021, 8, 663838. [Google Scholar] [CrossRef] [PubMed]
  78. Jheng, H.F.; Tsai, P.J.; Guo, S.M.; Kuo, L.H.; Chang, C.S.; Su, I.J.; Chang, C.R.; Tsai, Y.S. Mitochondrial fission contributes to mitochondrial dysfunction and insulin resistance in skeletal muscle. Mol. Cell. Biol. 2012, 32, 309–319. [Google Scholar] [CrossRef] [PubMed]
  79. Lambertucci, R.H.; Hirabara, S.M.; Silveira Ldos, R.; Levada-Pires, A.C.; Curi, R.; Pithon-Curi, T.C. Palmitate increases superoxide production through mitochondrial electron transport chain and NADPH oxidase activity in skeletal muscle cells. J. Cell. Physiol. 2008, 216, 796–804. [Google Scholar] [CrossRef]
  80. Livingston, M.J.; Wang, J.; Zhou, J.; Wu, G.; Ganley, I.G.; Hill, J.A.; Yin, X.M.; Dong, Z. Clearance of damaged mitochondria via mitophagy is important to the protective effect of ischemic preconditioning in kidneys. Autophagy 2019, 15, 2142–2162. [Google Scholar] [CrossRef]
  81. Feng, X.T.; Wang, T.Z.; Leng, J.; Chen, Y.; Liu, J.B.; Liu, Y.; Wang, W.J. Palmitate contributes to insulin resistance through downregulation of the Src-mediated phosphorylation of Akt in C2C12 myotubes. Biosci. Biotechnol. Biochem. 2012, 76, 1356–1361. [Google Scholar] [CrossRef]
  82. Burman, J.L.; Pickles, S.; Wang, C.; Sekine, S.; Vargas, J.N.S.; Zhang, Z.; Youle, A.M.; Nezich, C.L.; Wu, X.; Hammer, J.A.; et al. Mitochondrial fission facilitates the selective mitophagy of protein aggregates. J. Cell Biol. 2017, 216, 3231–3247. [Google Scholar] [CrossRef] [PubMed]
  83. Twig, G.; Elorza, A.; Molina, A.J.; Mohamed, H.; Wikstrom, J.D.; Walzer, G.; Stiles, L.; Haigh, S.E.; Katz, S.; Las, G.; et al. Fission and selective fusion govern mitochondrial segregation and elimination by autophagy. EMBO J. 2008, 27, 433–446. [Google Scholar] [CrossRef]
  84. Oh, S.J.; Park, K.; Sonn, S.K.; Oh, G.T.; Lee, M.S. Pancreatic beta-cell mitophagy as an adaptive response to metabolic stress and the underlying mechanism that involves lysosomal Ca2+ release. Exp. Mol. Med. 2023, 55, 1922–1932. [Google Scholar] [CrossRef] [PubMed]
  85. Geisler, S.; Holmstrom, K.M.; Skujat, D.; Fiesel, F.C.; Rothfuss, O.C.; Kahle, P.J.; Springer, W. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat. Cell Biol. 2010, 12, 119–131. [Google Scholar] [CrossRef] [PubMed]
  86. Ordureau, A.; Heo, J.M.; Duda, D.M.; Paulo, J.A.; Olszewski, J.L.; Yanishevski, D.; Rinehart, J.; Schulman, B.A.; Harper, J.W. Defining roles of PARKIN and ubiquitin phosphorylation by PINK1 in mitochondrial quality control using a ubiquitin replacement strategy. Proc. Natl. Acad. Sci. USA 2015, 112, 6637–6642. [Google Scholar] [CrossRef] [PubMed]
  87. Gladkova, C.; Maslen, S.L.; Skehel, J.M.; Komander, D. Mechanism of parkin activation by PINK1. Nature 2018, 559, 410–414. [Google Scholar] [CrossRef] [PubMed]
  88. Kane, L.A.; Lazarou, M.; Fogel, A.I.; Li, Y.; Yamano, K.; Sarraf, S.A.; Banerjee, S.; Youle, R.J. PINK1 phosphorylates ubiquitin to activate Parkin E3 ubiquitin ligase activity. J. Cell Biol. 2014, 205, 143–153. [Google Scholar] [CrossRef]
  89. Okatsu, K.; Oka, T.; Iguchi, M.; Imamura, K.; Kosako, H.; Tani, N.; Kimura, M.; Go, E.; Koyano, F.; Funayama, M.; et al. PINK1 autophosphorylation upon membrane potential dissipation is essential for Parkin recruitment to damaged mitochondria. Nat. Commun. 2012, 3, 1016. [Google Scholar] [CrossRef]
  90. Gao, A.; Jiang, J.; Xie, F.; Chen, L. Bnip3 in mitophagy: Novel insights and potential therapeutic target for diseases of secondary mitochondrial dysfunction. Clin. Chim. Acta 2020, 506, 72–83. [Google Scholar] [CrossRef]
  91. Tong, B.; Zhang, Z.; Li, X.; Liu, J.; Wang, H.; Song, L.; Feng, J.; Dai, Z.; Xu, Y. FUNDC1 modulates mitochondrial defects and pancreatic beta-cell dysfunction under lipotoxicity. Biochem. Biophys. Res. Commun. 2023, 672, 54–64. [Google Scholar] [CrossRef]
  92. Okatsu, K.; Kimura, M.; Oka, T.; Tanaka, K.; Matsuda, N. Unconventional PINK1 localization to the outer membrane of depolarized mitochondria drives Parkin recruitment. J. Cell Sci. 2015, 128, 964–978. [Google Scholar] [CrossRef] [PubMed]
  93. Greene, A.W.; Grenier, K.; Aguileta, M.A.; Muise, S.; Farazifard, R.; Haque, M.E.; McBride, H.M.; Park, D.S.; Fon, E.A. Mitochondrial processing peptidase regulates PINK1 processing, import and Parkin recruitment. EMBO Rep. 2012, 13, 378–385. [Google Scholar] [CrossRef] [PubMed]
  94. Aerts, L.; Craessaerts, K.; De Strooper, B.; Morais, V.A. PINK1 kinase catalytic activity is regulated by phosphorylation on serines 228 and 402. J. Biol. Chem. 2015, 290, 2798–2811. [Google Scholar] [CrossRef]
  95. Wauer, T.; Swatek, K.N.; Wagstaff, J.L.; Gladkova, C.; Pruneda, J.N.; Michel, M.A.; Gersch, M.; Johnson, C.M.; Freund, S.M.; Komander, D. Ubiquitin Ser65 phosphorylation affects ubiquitin structure, chain assembly and hydrolysis. EMBO J. 2015, 34, 307–325. [Google Scholar] [CrossRef] [PubMed]
  96. Wong, Y.C.; Holzbaur, E.L. Optineurin is an autophagy receptor for damaged mitochondria in parkin-mediated mitophagy that is disrupted by an ALS-linked mutation. Proc. Natl. Acad. Sci. USA 2014, 111, E4439–E4448. [Google Scholar] [CrossRef] [PubMed]
  97. Wong, Y.C.; Holzbaur, E.L. Temporal dynamics of PARK2/parkin and OPTN/optineurin recruitment during the mitophagy of damaged mitochondria. Autophagy 2015, 11, 422–424. [Google Scholar] [CrossRef] [PubMed]
  98. Kuang, Y.; Ma, K.; Zhou, C.; Ding, P.; Zhu, Y.; Chen, Q.; Xia, B. Structural basis for the phosphorylation of FUNDC1 LIR as a molecular switch of mitophagy. Autophagy 2016, 12, 2363–2373. [Google Scholar] [CrossRef]
  99. Wirth, M.; Zhang, W.; Razi, M.; Nyoni, L.; Joshi, D.; O’Reilly, N.; Johansen, T.; Tooze, S.A.; Mouilleron, S. Molecular determinants regulating selective binding of autophagy adapters and receptors to ATG8 proteins. Nat. Commun. 2019, 10, 2055. [Google Scholar] [CrossRef]
  100. Wei, Y.; Chiang, W.C.; Sumpter, R., Jr.; Mishra, P.; Levine, B. Prohibitin 2 Is an Inner Mitochondrial Membrane Mitophagy Receptor. Cell 2017, 168, 224–238.e10. [Google Scholar] [CrossRef]
  101. da Silva Rosa, S.C.; Martens, M.D.; Field, J.T.; Nguyen, L.; Kereliuk, S.M.; Hai, Y.; Chapman, D.; Diehl-Jones, W.; Aliani, M.; West, A.R.; et al. BNIP3L/Nix-induced mitochondrial fission, mitophagy, and impaired myocyte glucose uptake are abrogated by PRKA/PKA phosphorylation. Autophagy 2021, 17, 2257–2272. [Google Scholar] [CrossRef]
  102. Wu, H.; Wang, Y.; Li, W.; Chen, H.; Du, L.; Liu, D.; Wang, X.; Xu, T.; Liu, L.; Chen, Q. Deficiency of mitophagy receptor FUNDC1 impairs mitochondrial quality and aggravates dietary-induced obesity and metabolic syndrome. Autophagy 2019, 15, 1882–1898. [Google Scholar] [CrossRef]
  103. Fu, T.; Xu, Z.; Liu, L.; Guo, Q.; Wu, H.; Liang, X.; Zhou, D.; Xiao, L.; Liu, L.; Liu, Y.; et al. Mitophagy Directs Muscle-Adipose Crosstalk to Alleviate Dietary Obesity. Cell Rep. 2018, 23, 1357–1372. [Google Scholar] [CrossRef] [PubMed]
  104. Yan, C.; Gong, L.; Chen, L.; Xu, M.; Abou-Hamdan, H.; Tang, M.; Desaubry, L.; Song, Z. PHB2 (prohibitin 2) promotes PINK1-PRKN/Parkin-dependent mitophagy by the PARL-PGAM5-PINK1 axis. Autophagy 2020, 16, 419–434. [Google Scholar] [CrossRef] [PubMed]
  105. Spurlin, B.A.; Park, S.Y.; Nevins, A.K.; Kim, J.K.; Thurmond, D.C. Syntaxin 4 transgenic mice exhibit enhanced insulin-mediated glucose uptake in skeletal muscle. Diabetes 2004, 53, 2223–2231. [Google Scholar] [CrossRef] [PubMed]
  106. Yang, C.; Coker, K.J.; Kim, J.K.; Mora, S.; Thurmond, D.C.; Davis, A.C.; Yang, B.; Williamson, R.A.; Shulman, G.I.; Pessin, J.E. Syntaxin 4 heterozygous knockout mice develop muscle insulin resistance. J. Clin. Investig. 2001, 107, 1311–1318. [Google Scholar] [CrossRef] [PubMed]
  107. Kageyama, Y.; Hoshijima, M.; Seo, K.; Bedja, D.; Sysa-Shah, P.; Andrabi, S.A.; Chen, W.; Höke, A.; Dawson, V.L.; Dawson, T.M.; et al. Parkin-independent mitophagy requires Drp1 and maintains the integrity of mammalian heart and brain. EMBO J. 2014, 33, 2798–2813. [Google Scholar] [CrossRef] [PubMed]
  108. Pernaute, B.; Pérez-Montero, S.; Sánchez Nieto, J.M.; Di Gregorio, A.; Lima, A.; Lawlor, K.; Bowling, S.; Liccardi, G.; Tomás, A.; Meier, P.; et al. DRP1 levels determine the apoptotic threshold during embryonic differentiation through a mitophagy-dependent mechanism. Dev. Cell 2022, 57, 1316–1330.e7. [Google Scholar] [CrossRef] [PubMed]
  109. Panigrahi, D.P.; Praharaj, P.P.; Behera, B.P.; Patra, S.; Patil, S.; Patro, B.S.; Bhutia, S.K. The inner mitochondrial membrane fission protein MTP18 serves as a mitophagy receptor to prevent apoptosis in oral cancer. J. Cell Sci. 2023, 136, jcs259986. [Google Scholar] [CrossRef]
  110. Ehrlicher, S.E.; Stierwalt, H.D.; Newsom, S.A.; Robinson, M.M. Short-Term High-Fat Feeding Does Not Alter Mitochondrial Lipid Respiratory Capacity but Triggers Mitophagy Response in Skeletal Muscle of Mice. Front. Endocrinol. 2021, 12, 651211. [Google Scholar] [CrossRef]
  111. Lee, H.; Ha, T.Y.; Jung, C.H.; Nirmala, F.S.; Park, S.Y.; Huh, Y.H.; Ahn, J. Mitochondrial dysfunction in skeletal muscle contributes to the development of acute insulin resistance in mice. J. Cachexia Sarcopenia Muscle 2021, 12, 1925–1939. [Google Scholar] [CrossRef]
  112. Yeo, D.; Zhang, T.; Liu, T.; Zhang, Y.; Kang, C.; Ji, L.L. Protective effects of extra virgin olive oil and exercise training on rat skeletal muscle against high-fat diet feeding. J. Nutr. Biochem. 2022, 100, 108902. [Google Scholar] [CrossRef]
  113. Hyatt, J.K.; de Cabo, R.; Mattison, J.A. Resveratrol Blunts Mitochondrial Loss in Slow and Mixed Skeletal Muscle Phenotypes of Non-Human Primates following a Long-Term High Fat/Sugar Diet. J. Diet. Suppl. 2023, 20, 563–581. [Google Scholar] [CrossRef]
  114. Bhansali, S.; Bhansali, A.; Walia, R.; Saikia, U.N.; Dhawan, V. Alterations in Mitochondrial Oxidative Stress and Mitophagy in Subjects with Prediabetes and Type 2 Diabetes Mellitus. Front. Endocrinol. 2017, 8, 347. [Google Scholar] [CrossRef]
  115. Ohman, T.; Teppo, J.; Datta, N.; Makinen, S.; Varjosalo, M.; Koistinen, H.A. Skeletal muscle proteomes reveal downregulation of mitochondrial proteins in transition from prediabetes into type 2 diabetes. iScience 2021, 24, 102712. [Google Scholar] [CrossRef] [PubMed]
  116. Zhang, D.; Lee, J.H.; Shin, H.E.; Kwak, S.E.; Bae, J.H.; Tang, L.; Song, W. The Effects of Exercise and Restriction of Sugar-Sweetened Beverages on Muscle Function and Autophagy Regulation in High-Fat High-Sucrose-Fed Obesity Mice. Diabetes Metab. J. 2021, 45, 773–786. [Google Scholar] [CrossRef]
  117. Yeon, M.H.; Seo, E.; Lee, J.H.; Jun, H.S. Bavachin and Corylifol A Improve Muscle Atrophy by Enhancing Mitochondria Quality Control in Type 2 Diabetic Mice. Antioxidants 2023, 12, 137. [Google Scholar] [CrossRef] [PubMed]
  118. Chen, X.; Yi, L.; Song, S.; Wang, L.; Liang, Q.; Wang, Y.; Wu, Y.; Gao, Q. Puerarin attenuates palmitate-induced mitochondrial dysfunction, impaired mitophagy and inflammation in L6 myotubes. Life Sci. 2018, 206, 84–92. [Google Scholar] [CrossRef] [PubMed]
  119. Mughal, W.; Nguyen, L.; Pustylnik, S.; da Silva Rosa, S.C.; Piotrowski, S.; Chapman, D.; Du, M.; Alli, N.S.; Grigull, J.; Halayko, A.J.; et al. A conserved MADS-box phosphorylation motif regulates differentiation and mitochondrial function in skeletal, cardiac, and smooth muscle cells. Cell Death Dis. 2015, 6, e1944. [Google Scholar] [CrossRef] [PubMed]
  120. Almeida, E.A.; Mehndiratta, M.; Madhu, S.V.; Kar, R.; Puri, D. PINK1 and oxidative stress in lean and obese patients with type 2 diabetes mellitus. J. Diabetes Complicat. 2023, 37, 108542. [Google Scholar] [CrossRef] [PubMed]
  121. Merz, K.E.; Tunduguru, R.; Ahn, M.; Salunkhe, V.A.; Veluthakal, R.; Hwang, J.; Bhattacharya, S.; McCown, E.M.; Garcia, P.A.; Zhou, C.; et al. Changes in Skeletal Muscle PAK1 Levels Regulate Tissue Crosstalk to Impact Whole Body Glucose Homeostasis. Front. Endocrinol. 2022, 13, 821849. [Google Scholar] [CrossRef]
  122. Balakrishnan, R.G.P.; Veluthakal, R.; Huss, J.M.; Hoolachan, J.M.; Thurmond, D.C. Toward Ameliorating Insulin Resistance: Targeting a Novel PAK1 Signaling Pathway Required for Skeletal Muscle Mitochondrial Function. Antioxidants 2023, 12, 1658. [Google Scholar] [CrossRef] [PubMed]
  123. Ahn, M.; Oh, E.; McCown, E.M.; Wang, X.; Veluthakal, R.; Thurmond, D.C. A requirement for PAK1 to support mitochondrial function and maintain cellular redox balance via electron transport chain proteins to prevent β-cell apoptosis. Metabolism 2021, 115, 154431. [Google Scholar] [CrossRef] [PubMed]
  124. Gallagher, E.J.; LeRoith, D.; Stasinopoulos, M.; Zelenko, Z.; Shiloach, J. Polyol accumulation in muscle and liver in a mouse model of type 2 diabetes. J. Diabetes Complicat. 2016, 30, 999–1007. [Google Scholar] [CrossRef] [PubMed]
  125. Du, H.; Ma, Y.; Wang, X.; Zhang, Y.; Zhu, L.; Shi, S.; Pan, S.; Liu, Z. Advanced glycation end products induce skeletal muscle atrophy and insulin resistance via activating ROS-mediated ER stress PERK/FOXO1 signaling. Am. J. Physiol. Endocrinol. Metab. 2023, 324, E279–E287. [Google Scholar] [CrossRef] [PubMed]
  126. Davis, K.E.; Prasad, C.; Vijayagopal, P.; Juma, S.; Imrhan, V. Advanced Glycation End Products, Inflammation, and Chronic Metabolic Diseases: Links in a Chain? Crit. Rev. Food Sci. Nutr. 2016, 56, 989–998. [Google Scholar] [CrossRef]
  127. Wu, Y.; Song, P.; Xu, J.; Zhang, M.; Zou, M.H. Activation of protein phosphatase 2A by palmitate inhibits AMP-activated protein kinase. J. Biol. Chem. 2007, 282, 9777–9788. [Google Scholar] [CrossRef] [PubMed]
  128. Chung, S.S.; Ho, E.C.; Lam, K.S.; Chung, S.K. Contribution of polyol pathway to diabetes-induced oxidative stress. J. Am. Soc. Nephrol. 2003, 14, S233–S236. [Google Scholar] [CrossRef] [PubMed]
  129. Yan, L.J. Pathogenesis of chronic hyperglycemia: From reductive stress to oxidative stress. J. Diabetes Res. 2014, 2014, 137919. [Google Scholar] [CrossRef]
  130. Andrich, D.E.; Melbouci, L.; Ou, Y.; Auclair, N.; Mercier, J.; Grenier, J.C.; Lira, F.S.; Barreiro, L.B.; Danialou, G.; Comtois, A.S.; et al. A Short-Term High-Fat Diet Alters Glutathione Levels and IL-6 Gene Expression in Oxidative Skeletal Muscles of Young Rats. Front. Physiol. 2019, 10, 372. [Google Scholar] [CrossRef]
  131. Sondergard, S.D.; Cintin, I.; Kuhlman, A.B.; Morville, T.H.; Bergmann, M.L.; Kjaer, L.K.; Poulsen, H.E.; Giustarini, D.; Rossi, R.; Dela, F.; et al. The effects of 3 weeks of oral glutathione supplementation on whole body insulin sensitivity in obese males with and without type 2 diabetes: A randomized trial. Appl. Physiol. Nutr. Metab. 2021, 46, 1133–1142. [Google Scholar] [CrossRef]
  132. Patti, M.E.; Virkamaki, A.; Landaker, E.J.; Kahn, C.R.; Yki-Jarvinen, H. Activation of the hexosamine pathway by glucosamine in vivo induces insulin resistance of early postreceptor insulin signaling events in skeletal muscle. Diabetes 1999, 48, 1562–1571. [Google Scholar] [CrossRef] [PubMed]
  133. Buse, M.G.; Robinson, K.A.; Marshall, B.A.; Mueckler, M. Differential effects of GLUT1 or GLUT4 overexpression on hexosamine biosynthesis by muscles of transgenic mice. J. Biol. Chem. 1996, 271, 23197–23202. [Google Scholar] [CrossRef] [PubMed]
  134. Habegger, K.M.; Penque, B.A.; Sealls, W.; Tackett, L.; Bell, L.N.; Blue, E.K.; Gallagher, P.J.; Sturek, M.; Alloosh, M.A.; Steinberg, H.O.; et al. Fat-induced membrane cholesterol accrual provokes cortical filamentous actin destabilisation and glucose transport dysfunction in skeletal muscle. Diabetologia 2012, 55, 457–467. [Google Scholar] [CrossRef] [PubMed]
  135. Penque, B.A.; Hoggatt, A.M.; Herring, B.P.; Elmendorf, J.S. Hexosamine biosynthesis impairs insulin action via a cholesterolgenic response. Mol. Endocrinol. 2013, 27, 536–547. [Google Scholar] [CrossRef] [PubMed]
  136. Bhonagiri, P.; Pattar, G.R.; Habegger, K.M.; McCarthy, A.M.; Tackett, L.; Elmendorf, J.S. Evidence coupling increased hexosamine biosynthesis pathway activity to membrane cholesterol toxicity and cortical filamentous actin derangement contributing to cellular insulin resistance. Endocrinology 2011, 152, 3373–3384. [Google Scholar] [CrossRef] [PubMed]
  137. Haga, Y.; Ishii, K.; Suzuki, T. N-glycosylation is critical for the stability and intracellular trafficking of glucose transporter GLUT4. J. Biol. Chem. 2011, 286, 31320–31327. [Google Scholar] [CrossRef] [PubMed]
  138. Yki-Jarvinen, H.; Daniels, M.C.; Virkamaki, A.; Makimattila, S.; DeFronzo, R.A.; McClain, D. Increased glutamine:fructose-6-phosphate amidotransferase activity in skeletal muscle of patients with NIDDM. Diabetes 1996, 45, 302–307. [Google Scholar] [CrossRef] [PubMed]
  139. Srinivasan, V.; Sandhya, N.; Sampathkumar, R.; Farooq, S.; Mohan, V.; Balasubramanyam, M. Glutamine fructose-6-phosphate amidotransferase (GFAT) gene expression and activity in patients with type 2 diabetes: Inter-relationships with hyperglycaemia and oxidative stress. Clin. Biochem. 2007, 40, 952–957. [Google Scholar] [CrossRef]
  140. Coughlan, M.T.; Thorburn, D.R.; Penfold, S.A.; Laskowski, A.; Harcourt, B.E.; Sourris, K.C.; Tan, A.L.; Fukami, K.; Thallas-Bonke, V.; Nawroth, P.P.; et al. RAGE-induced cytosolic ROS promote mitochondrial superoxide generation in diabetes. J. Am. Soc. Nephrol. 2009, 20, 742–752. [Google Scholar] [CrossRef]
  141. Xing, Y.; Pan, S.; Zhu, L.; Cui, Q.; Tang, Z.; Liu, Z.; Liu, F. Advanced Glycation End Products Induce Atherosclerosis via RAGE/TLR4 Signaling Mediated-M1 Macrophage Polarization-Dependent Vascular Smooth Muscle Cell Phenotypic Conversion. Oxid. Med. Cell Longev. 2022, 2022, 9763377. [Google Scholar] [CrossRef]
  142. Chen, Y.H.; Chen, Z.W.; Li, H.M.; Yan, X.F.; Feng, B. AGE/RAGE-Induced EMP Release via the NOX-Derived ROS Pathway. J. Diabetes Res. 2018, 2018, 6823058. [Google Scholar] [CrossRef] [PubMed]
  143. Hanssen, N.M.; Beulens, J.W.; van Dieren, S.; Scheijen, J.L.; van der Daphne, A.; Spijkerman, A.M.; van der Schouw, Y.T.; Stehouwer, C.D.; Schalkwijk, C.G. Plasma advanced glycation end products are associated with incident cardiovascular events in individuals with type 2 diabetes: A case-cohort study with a median follow-up of 10 years (EPIC-NL). Diabetes 2015, 64, 257–265. [Google Scholar] [CrossRef] [PubMed]
  144. Cassese, A.; Esposito, I.; Fiory, F.; Barbagallo, A.P.; Paturzo, F.; Mirra, P.; Ulianich, L.; Giacco, F.; Iadicicco, C.; Lombardi, A.; et al. In skeletal muscle advanced glycation end products (AGEs) inhibit insulin action and induce the formation of multimolecular complexes including the receptor for AGEs. J. Biol. Chem. 2008, 283, 36088–36099. [Google Scholar] [CrossRef] [PubMed]
  145. Chiappalupi, S.; Sorci, G.; Vukasinovic, A.; Salvadori, L.; Sagheddu, R.; Coletti, D.; Renga, G.; Romani, L.; Donato, R.; Riuzzi, F. Targeting RAGE prevents muscle wasting and prolongs survival in cancer cachexia. J. Cachexia Sarcopenia Muscle 2020, 11, 929–946. [Google Scholar] [CrossRef] [PubMed]
  146. Mori, H.; Kuroda, A.; Ishizu, M.; Ohishi, M.; Takashi, Y.; Otsuka, Y.; Taniguchi, S.; Tamaki, M.; Kurahashi, K.; Yoshida, S.; et al. Association of accumulated advanced glycation end-products with a high prevalence of sarcopenia and dynapenia in patients with type 2 diabetes. J. Diabetes Investig. 2019, 10, 1332–1340. [Google Scholar] [CrossRef] [PubMed]
  147. Weir, G.C. Glucolipotoxicity, beta-Cells, and Diabetes: The Emperor Has No Clothes. Diabetes 2020, 69, 273–278. [Google Scholar] [CrossRef] [PubMed]
  148. Hall, E.; Jonsson, J.; Ofori, J.K.; Volkov, P.; Perfilyev, A.; Dekker Nitert, M.; Eliasson, L.; Ling, C.; Bacos, K. Glucolipotoxicity Alters Insulin Secretion via Epigenetic Changes in Human Islets. Diabetes 2019, 68, 1965–1974. [Google Scholar] [CrossRef] [PubMed]
  149. Brun, T.; Li, N.; Jourdain, A.A.; Gaudet, P.; Duhamel, D.; Meyer, J.; Bosco, D.; Maechler, P. Diabetogenic milieus induce specific changes in mitochondrial transcriptome and differentiation of human pancreatic islets. Hum. Mol. Genet. 2015, 24, 5270–5284. [Google Scholar] [CrossRef]
  150. Tiedge, M.; Lortz, S.; Drinkgern, J.; Lenzen, S. Relation between antioxidant enzyme gene expression and antioxidative defense status of insulin-producing cells. Diabetes 1997, 46, 1733–1742. [Google Scholar] [CrossRef]
  151. Lenzen, S.; Drinkgern, J.; Tiedge, M. Low antioxidant enzyme gene expression in pancreatic islets compared with various other mouse tissues. Free Radic. Biol. Med. 1996, 20, 463–466. [Google Scholar] [CrossRef]
  152. Argaev-Frenkel, L.; Rosenzweig, T. Redox Balance in Type 2 Diabetes: Therapeutic Potential and the Challenge of Antioxidant-Based Therapy. Antioxidants 2023, 12, 994. [Google Scholar] [CrossRef] [PubMed]
  153. Stancill, J.S.; Corbett, J.A. Hydrogen peroxide detoxification through the peroxiredoxin/thioredoxin antioxidant system: A look at the pancreatic beta-cell oxidant defense. Vitam. Horm. 2023, 121, 45–66. [Google Scholar] [CrossRef] [PubMed]
  154. Zengi, A.; Ercan, G.; Caglayan, O.; Tamsel, S.; Karadeniz, M.; Simsir, I.; Harman, E.; Kahraman, C.; Orman, M.; Cetinkalp, S.; et al. Increased oxidative DNA damage in lean normoglycemic offspring of type 2 diabetic patients. Exp. Clin. Endocrinol. Diabetes 2011, 119, 467–471. [Google Scholar] [CrossRef] [PubMed]
  155. Okamoto, K.; Iwasaki, N.; Nishimura, C.; Doi, K.; Noiri, E.; Nakamura, S.; Takizawa, M.; Ogata, M.; Fujimaki, R.; Grarup, N.; et al. Identification of KCNJ15 as a susceptibility gene in Asian patients with type 2 diabetes mellitus. Am. J. Hum. Genet. 2010, 86, 54–64. [Google Scholar] [CrossRef] [PubMed]
  156. Wang, Z.; Oh, E.; Clapp, D.W.; Chernoff, J.; Thurmond, D.C. Inhibition or ablation of p21-activated kinase (PAK1) disrupts glucose homeostatic mechanisms in vivo. J. Biol. Chem. 2011, 286, 41359–41367. [Google Scholar] [CrossRef] [PubMed]
  157. Yoder, S.M.; Dineen, S.L.; Wang, Z.; Thurmond, D.C. YES, a Src family kinase, is a proximal glucose-specific activator of cell division cycle control protein 42 (Cdc42) in pancreatic islet β cells. J. Biol. Chem. 2014, 289, 11476–11487. [Google Scholar] [CrossRef]
  158. Janssen, R.J.; Distelmaier, F.; Smeets, R.; Wijnhoven, T.; Østergaard, E.; Jaspers, N.G.; Raams, A.; Kemp, S.; Rodenburg, R.J.; Willems, P.H.; et al. Contiguous gene deletion of ELOVL7, ERCC8 and NDUFAF2 in a patient with a fatal multisystem disorder. Hum. Mol. Genet. 2009, 18, 3365–3374. [Google Scholar] [CrossRef]
  159. Anello, M.; Lupi, R.; Spampinato, D.; Piro, S.; Masini, M.; Boggi, U.; Del Prato, S.; Rabuazzo, A.M.; Purrello, F.; Marchetti, P. Functional and morphological alterations of mitochondria in pancreatic beta cells from type 2 diabetic patients. Diabetologia 2005, 48, 282–289. [Google Scholar] [CrossRef]
  160. Maechler, P.; Wollheim, C.B. Mitochondrial function in normal and diabetic beta-cells. Nature 2001, 414, 807–812. [Google Scholar] [CrossRef]
  161. Deng, S.; Vatamaniuk, M.; Huang, X.; Doliba, N.; Lian, M.M.; Frank, A.; Velidedeoglu, E.; Desai, N.M.; Koeberlein, B.; Wolf, B.; et al. Structural and functional abnormalities in the islets isolated from type 2 diabetic subjects. Diabetes 2004, 53, 624–632. [Google Scholar] [CrossRef]
  162. Weksler-Zangen, S. Is Type 2 Diabetes a Primary Mitochondrial Disorder? Cells 2022, 11, 1617. [Google Scholar] [CrossRef]
  163. Mulder, H. Transcribing beta-cell mitochondria in health and disease. Mol. Metab. 2017, 6, 1040–1051. [Google Scholar] [CrossRef]
  164. Molina, A.J.; Wikstrom, J.D.; Stiles, L.; Las, G.; Mohamed, H.; Elorza, A.; Walzer, G.; Twig, G.; Katz, S.; Corkey, B.E.; et al. Mitochondrial networking protects beta-cells from nutrient-induced apoptosis. Diabetes 2009, 58, 2303–2315. [Google Scholar] [CrossRef]
  165. Dlaskova, A.; Spacek, T.; Santorova, J.; Plecita-Hlavata, L.; Berkova, Z.; Saudek, F.; Lessard, M.; Bewersdorf, J.; Jezek, P. 4Pi microscopy reveals an impaired three-dimensional mitochondrial network of pancreatic islet beta-cells, an experimental model of type-2 diabetes. Biochim. Biophys. Acta 2010, 1797, 1327–1341. [Google Scholar] [CrossRef] [PubMed]
  166. Shimabukuro, M.; Zhou, Y.T.; Levi, M.; Unger, R.H. Fatty acid-induced beta cell apoptosis: A link between obesity and diabetes. Proc. Natl. Acad. Sci. USA 1998, 95, 2498–2502. [Google Scholar] [CrossRef] [PubMed]
  167. Lee, J.H.; Mellado-Gil, J.M.; Bahn, Y.J.; Pathy, S.M.; Zhang, Y.E.; Rane, S.G. Protection from beta-cell apoptosis by inhibition of TGF-beta/Smad3 signaling. Cell Death Dis. 2020, 11, 184. [Google Scholar] [CrossRef] [PubMed]
  168. Weksler-Zangen, S.; Raz, I.; Lenzen, S.; Jorns, A.; Ehrenfeld, S.; Amir, G.; Oprescu, A.; Yagil, Y.; Yagil, C.; Zangen, D.H.; et al. Impaired glucose-stimulated insulin secretion is coupled with exocrine pancreatic lesions in the Cohen diabetic rat. Diabetes 2008, 57, 279–287. [Google Scholar] [CrossRef] [PubMed]
  169. Aharon-Hananel, G.; Romero-Afrima, L.; Saada, A.; Mantzur, C.; Raz, I.; Weksler-Zangen, S. Cytochrome c Oxidase Activity as a Metabolic Regulator in Pancreatic Beta-Cells. Cells 2022, 11, 929. [Google Scholar] [CrossRef] [PubMed]
  170. Weksler-Zangen, S.; Aharon-Hananel, G.; Mantzur, C.; Aouizerat, T.; Gurgul-Convey, E.; Raz, I.; Saada, A. IL-1beta hampers glucose-stimulated insulin secretion in Cohen diabetic rat islets through mitochondrial cytochrome c oxidase inhibition by nitric oxide. Am. J. Physiol. Endocrinol. Metab. 2014, 306, E648–E657. [Google Scholar] [CrossRef] [PubMed]
  171. Sidarala, V.; Pearson, G.L.; Parekh, V.S.; Thompson, B.; Christen, L.; Gingerich, M.A.; Zhu, J.; Stromer, T.; Ren, J.; Reck, E.C.; et al. Mitophagy protects beta cells from inflammatory damage in diabetes. JCI Insight 2020, 5, e141138. [Google Scholar] [CrossRef]
  172. Guo, T.; Liu, T.; Sun, Y.; Liu, X.; Xiong, R.; Li, H.; Li, Z.; Zhang, Z.; Tian, Z.; Tian, Y. Sonodynamic therapy inhibits palmitate-induced beta cell dysfunction via PINK1/Parkin-dependent mitophagy. Cell Death Dis. 2019, 10, 457. [Google Scholar] [CrossRef] [PubMed]
  173. Huang, Q.; Liu, Z.; Yang, Y.; Yang, Y.; Huang, T.; Hong, Y.; Zhang, J.; Chen, Q.; Zhao, T.; Xiao, Z.; et al. Selenium Nanodots (SENDs) as Antioxidants and Antioxidant-Prodrugs to Rescue Islet beta Cells in Type 2 Diabetes Mellitus by Restoring Mitophagy and Alleviating Endoplasmic Reticulum Stress. Adv. Sci. 2023, 10, e2300880. [Google Scholar] [CrossRef] [PubMed]
  174. Rieusset, J. The role of endoplasmic reticulum-mitochondria contact sites in the control of glucose homeostasis: An update. Cell Death Dis. 2018, 9, 388. [Google Scholar] [CrossRef]
  175. Thivolet, C.; Vial, G.; Cassel, R.; Rieusset, J.; Madec, A.M. Reduction of endoplasmic reticulum-mitochondria interactions in beta cells from patients with type 2 diabetes. PLoS ONE 2017, 12, e0182027. [Google Scholar] [CrossRef] [PubMed]
  176. Herrera-Cruz, M.S.; Simmen, T. Over Six Decades of Discovery and Characterization of the Architecture at Mitochondria-Associated Membranes (MAMs). Adv. Exp. Med. Biol. 2017, 997, 13–31. [Google Scholar] [CrossRef] [PubMed]
  177. Csordas, G.; Varnai, P.; Golenar, T.; Roy, S.; Purkins, G.; Schneider, T.G.; Balla, T.; Hajnoczky, G. Imaging interorganelle contacts and local calcium dynamics at the ER-mitochondrial interface. Mol. Cell 2010, 39, 121–132. [Google Scholar] [CrossRef] [PubMed]
  178. Yang, S.; Zhou, R.; Zhang, C.; He, S.; Su, Z. Mitochondria-Associated Endoplasmic Reticulum Membranes in the Pathogenesis of Type 2 Diabetes Mellitus. Front. Cell Dev. Biol. 2020, 8, 571554. [Google Scholar] [CrossRef]
  179. Lupi, R.; Dotta, F.; Marselli, L.; Del Guerra, S.; Masini, M.; Santangelo, C.; Patane, G.; Boggi, U.; Piro, S.; Anello, M.; et al. Prolonged exposure to free fatty acids has cytostatic and pro-apoptotic effects on human pancreatic islets: Evidence that beta-cell death is caspase mediated, partially dependent on ceramide pathway, and Bcl-2 regulated. Diabetes 2002, 51, 1437–1442. [Google Scholar] [CrossRef]
  180. Veluthakal, R.; Palanivel, R.; Zhao, Y.; McDonald, P.; Gruber, S.; Kowluru, A. Ceramide induces mitochondrial abnormalities in insulin-secreting INS-1 cells: Potential mechanisms underlying ceramide-mediated metabolic dysfunction of the beta cell. Apoptosis 2005, 10, 841–850. [Google Scholar] [CrossRef]
  181. Thery, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): A position statement of the International Society for Extracellular Vesicles and update of the MISEV2014 guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef]
  182. Wang, J.; Li, L.; Zhang, Z.; Zhang, X.; Zhu, Y.; Zhang, C.; Bi, Y. Extracellular vesicles mediate the communication of adipose tissue with brain and promote cognitive impairment associated with insulin resistance. Cell Metab. 2022, 34, 1264–1279.e8. [Google Scholar] [CrossRef]
  183. Chidester, S.; Livinski, A.A.; Fish, A.F.; Joseph, P.V. The Role of Extracellular Vesicles in β-Cell Function and Viability: A Scoping Review. Front. Endocrinol. 2020, 11, 375. [Google Scholar] [CrossRef] [PubMed]
  184. Liu, J.; Sun, X.; Zhang, F.L.; Jin, H.; Yan, X.L.; Huang, S.; Guo, Z.N.; Yang, Y. Clinical Potential of Extracellular Vesicles in Type 2 Diabetes. Front. Endocrinol. 2020, 11, 596811. [Google Scholar] [CrossRef] [PubMed]
  185. Watanabe, S.; Sudo, Y.; Makino, T.; Kimura, S.; Tomita, K.; Noguchi, M.; Sakurai, H.; Shimizu, M.; Takahashi, Y.; Sato, R.; et al. Skeletal muscle releases extracellular vesicles with distinct protein and microRNA signatures that function in the muscle microenvironment. PNAS Nexus 2022, 1, pgac173. [Google Scholar] [CrossRef] [PubMed]
  186. Pedersen, B.K. Muscle as a secretory organ. Compr. Physiol. 2013, 3, 1337–1362. [Google Scholar] [CrossRef] [PubMed]
  187. Jalabert, A.; Vial, G.; Guay, C.; Wiklander, O.P.; Nordin, J.Z.; Aswad, H.; Forterre, A.; Meugnier, E.; Pesenti, S.; Regazzi, R.; et al. Exosome-like vesicles released from lipid-induced insulin-resistant muscles modulate gene expression and proliferation of beta recipient cells in mice. Diabetologia 2016, 59, 1049–1058. [Google Scholar] [CrossRef] [PubMed]
  188. Aswad, H.; Forterre, A.; Wiklander, O.P.; Vial, G.; Danty-Berger, E.; Jalabert, A.; Lamaziere, A.; Meugnier, E.; Pesenti, S.; Ott, C.; et al. Exosomes participate in the alteration of muscle homeostasis during lipid-induced insulin resistance in mice. Diabetologia 2014, 57, 2155–2164. [Google Scholar] [CrossRef] [PubMed]
  189. Carson, B.P. The Potential Role of Contraction-Induced Myokines in the Regulation of Metabolic Function for the Prevention and Treatment of Type 2 Diabetes. Front. Endocrinol. 2017, 8, 97. [Google Scholar] [CrossRef]
  190. Chen, X.; Wang, K.; Chen, J.; Guo, J.; Yin, Y.; Cai, X.; Guo, X.; Wang, G.; Yang, R.; Zhu, L.; et al. In vitro evidence suggests that miR-133a-mediated regulation of uncoupling protein 2 (UCP2) is an indispensable step in myogenic differentiation. J. Biol. Chem. 2009, 284, 5362–5369. [Google Scholar] [CrossRef]
  191. Robson-Doucette, C.A.; Sultan, S.; Allister, E.M.; Wikstrom, J.D.; Koshkin, V.; Bhattacharjee, A.; Prentice, K.J.; Sereda, S.B.; Shirihai, O.S.; Wheeler, M.B. Beta-cell uncoupling protein 2 regulates reactive oxygen species production, which influences both insulin and glucagon secretion. Diabetes 2011, 60, 2710–2719. [Google Scholar] [CrossRef]
  192. Lisi, V.; Senesi, G.; Balbi, C. Converging protective pathways: Exploring the linkage between physical exercise, extracellular vesicles and oxidative stress. Free Radic. Biol. Med. 2023, 208, 718–727. [Google Scholar] [CrossRef] [PubMed]
  193. Kobayashi, Y.; Eguchi, A.; Tamai, Y.; Fukuda, S.; Tempaku, M.; Izuoka, K.; Iwasa, M.; Takei, Y.; Togashi, K. Protein Composition of Circulating Extracellular Vesicles Immediately Changed by Particular Short Time of High-Intensity Interval Training Exercise. Front. Physiol. 2021, 12, 693007. [Google Scholar] [CrossRef] [PubMed]
  194. Plomgaard, P.; Halban, P.A.; Bouzakri, K. Bimodal impact of skeletal muscle on pancreatic beta-cell function in health and disease. Diabetes Obes. Metab. 2012, 14 (Suppl. 3), 78–84. [Google Scholar] [CrossRef] [PubMed]
  195. Natalicchio, A.; Marrano, N.; Biondi, G.; Spagnuolo, R.; Labarbuta, R.; Porreca, I.; Cignarelli, A.; Bugliani, M.; Marchetti, P.; Perrini, S.; et al. The Myokine Irisin Is Released in Response to Saturated Fatty Acids and Promotes Pancreatic beta-Cell Survival and Insulin Secretion. Diabetes 2017, 66, 2849–2856. [Google Scholar] [CrossRef] [PubMed]
  196. Lee, Y.S.; Morinaga, H.; Kim, J.J.; Lagakos, W.; Taylor, S.; Keshwani, M.; Perkins, G.; Dong, H.; Kayali, A.G.; Sweet, I.R.; et al. The fractalkine/CX3CR1 system regulates beta cell function and insulin secretion. Cell 2013, 153, 413–425. [Google Scholar] [CrossRef] [PubMed]
  197. Riopel, M.; Seo, J.B.; Bandyopadhyay, G.K.; Li, P.; Wollam, J.; Chung, H.; Jung, S.R.; Murphy, A.; Wilson, M.; de Jong, R.; et al. Chronic fractalkine administration improves glucose tolerance and pancreatic endocrine function. J. Clin. Investig. 2018, 128, 1458–1470. [Google Scholar] [CrossRef] [PubMed]
  198. Fred, R.G.; Bang-Berthelsen, C.H.; Mandrup-Poulsen, T.; Grunnet, L.G.; Welsh, N. High glucose suppresses human islet insulin biosynthesis by inducing miR-133a leading to decreased polypyrimidine tract binding protein-expression. PLoS ONE 2010, 5, e10843. [Google Scholar] [CrossRef] [PubMed]
  199. Bouzakri, K.; Plomgaard, P.; Berney, T.; Donath, M.Y.; Pedersen, B.K.; Halban, P.A. Bimodal effect on pancreatic beta-cells of secretory products from normal or insulin-resistant human skeletal muscle. Diabetes 2011, 60, 1111–1121. [Google Scholar] [CrossRef]
  200. Javeed, N.; Her, T.K.; Brown, M.R.; Vanderboom, P.; Rakshit, K.; Egan, A.M.; Vella, A.; Lanza, I.; Matveyenko, A.V. Pro-inflammatory beta cell small extracellular vesicles induce beta cell failure through activation of the CXCL10/CXCR3 axis in diabetes. Cell Rep. 2021, 36, 109613. [Google Scholar] [CrossRef]
  201. Clement, E.; Lazar, I.; Attane, C.; Carrie, L.; Dauvillier, S.; Ducoux-Petit, M.; Esteve, D.; Menneteau, T.; Moutahir, M.; Le Gonidec, S.; et al. Adipocyte extracellular vesicles carry enzymes and fatty acids that stimulate mitochondrial metabolism and remodeling in tumor cells. EMBO J. 2020, 39, e102525. [Google Scholar] [CrossRef]
  202. Crewe, C.; Funcke, J.B.; Li, S.; Joffin, N.; Gliniak, C.M.; Ghaben, A.L.; An, Y.A.; Sadek, H.A.; Gordillo, R.; Akgul, Y.; et al. Extracellular vesicle-based interorgan transport of mitochondria from energetically stressed adipocytes. Cell Metab. 2021, 33, 1853–1868.e11. [Google Scholar] [CrossRef] [PubMed]
  203. Rosina, M.; Ceci, V.; Turchi, R.; Chuan, L.; Borcherding, N.; Sciarretta, F.; Sanchez-Diaz, M.; Tortolici, F.; Karlinsey, K.; Chiurchiu, V.; et al. Ejection of damaged mitochondria and their removal by macrophages ensure efficient thermogenesis in brown adipose tissue. Cell Metab. 2022, 34, 533–548. [Google Scholar] [CrossRef] [PubMed]
  204. Rackham, C.L.; Hubber, E.L.; Czajka, A.; Malik, A.N.; King, A.J.F.; Jones, P.M. Optimizing beta cell function through mesenchymal stromal cell-mediated mitochondria transfer. Stem Cells 2020, 38, 574–584. [Google Scholar] [CrossRef] [PubMed]
  205. Zhao, H.; Shang, Q.; Pan, Z.; Bai, Y.; Li, Z.; Zhang, H.; Zhang, Q.; Guo, C.; Zhang, L.; Wang, Q. Exosomes from Adipose-Derived Stem Cells Attenuate Adipose Inflammation and Obesity Through Polarizing M2 Macrophages and Beiging in White Adipose Tissue. Diabetes 2018, 67, 235–247. [Google Scholar] [CrossRef] [PubMed]
  206. Crewe, C.; Chen, S.; Bu, D.; Gliniak, C.M.; Wernstedt Asterholm, I.; Yu, X.X.; Joffin, N.; de Souza, C.O.; Funcke, J.B.; Oh, D.Y.; et al. Deficient Caveolin-1 Synthesis in Adipocytes Stimulates Systemic Insulin-Independent Glucose Uptake via Extracellular Vesicles. Diabetes 2022, 71, 2496–2512. [Google Scholar] [CrossRef] [PubMed]
  207. Kulaj, K.; Harger, A.; Bauer, M.; Caliskan, O.S.; Gupta, T.K.; Chiang, D.M.; Milbank, E.; Reber, J.; Karlas, A.; Kotzbeck, P.; et al. Adipocyte-derived extracellular vesicles increase insulin secretion through transport of insulinotropic protein cargo. Nat. Commun. 2023, 14, 709. [Google Scholar] [CrossRef] [PubMed]
  208. Adiga, D.; Bhat, S.; Shukla, V.; Shah, H.V.; Kuthethur, R.; Chakrabarty, S.; Kabekkodu, S.P. Double C-2 like domain beta (DOC2B) induces calcium dependent oxidative stress to promote lipotoxicity and mitochondrial dysfunction for its tumor suppressive function. Free Radic. Biol. Med. 2023, 201, 1–13. [Google Scholar] [CrossRef] [PubMed]
  209. Esparza, D.; Oh, E.; Ghaeli, I.; Jovanovic-Talisman, T.; Thurmond, D.C. 1770-P: Islet ß-Cells Shed DOC2B-Laden EVs via Sorting Signals Present in DOC2B Protein. Diabetes 2023, 72, 1770. [Google Scholar] [CrossRef]
  210. He, F.; Huang, Y.; Song, Z.; Zhou, H.J.; Zhang, H.; Perry, R.J.; Shulman, G.I.; Min, W. Mitophagy-mediated adipose inflammation contributes to type 2 diabetes with hepatic insulin resistance. J. Exp. Med. 2021, 218, e20201416. [Google Scholar] [CrossRef]
  211. Zhu, J.; Wu, F.; Li, C.; Mao, J.; Wang, Y.; Zhou, X.; Xie, H.; Wen, C. Application of Single Extracellular Vesicle Analysis Techniques. Int. J. Nanomed. 2023, 18, 5365–5376. [Google Scholar] [CrossRef]
  212. Brestoff, J.R.; Wilen, C.B.; Moley, J.R.; Li, Y.; Zou, W.; Malvin, N.P.; Rowen, M.N.; Saunders, B.T.; Ma, H.; Mack, M.R.; et al. Intercellular Mitochondria Transfer to Macrophages Regulates White Adipose Tissue Homeostasis and Is Impaired in Obesity. Cell Metab. 2021, 33, 270–282.e278. [Google Scholar] [CrossRef] [PubMed]
  213. Deng, Z.B.; Poliakov, A.; Hardy, R.W.; Clements, R.; Liu, C.; Liu, Y.; Wang, J.; Xiang, X.; Zhang, S.; Zhuang, X.; et al. Adipose tissue exosome-like vesicles mediate activation of macrophage-induced insulin resistance. Diabetes 2009, 58, 2498–2505. [Google Scholar] [CrossRef] [PubMed]
  214. Sadeghi, A.; Rostamirad, A.; Seyyedebrahimi, S.; Meshkani, R. Curcumin ameliorates palmitate-induced inflammation in skeletal muscle cells by regulating JNK/NF-kB pathway and ROS production. Inflammopharmacology 2018, 26, 1265–1272. [Google Scholar] [CrossRef] [PubMed]
  215. Mahdavi, A.; Moradi, S.; Askari, G.; Iraj, B.; Sathyapalan, T.; Guest, P.C.; Bagherniya, M.; Sahebkar, A. Effect of Curcumin on Glycemic Control in Patients with Type 2 Diabetes: A Systematic Review of Randomized Clinical Trials. Adv. Exp. Med. Biol. 2021, 1291, 139–149. [Google Scholar] [CrossRef] [PubMed]
  216. Oliveira, S.; Monteiro-Alfredo, T.; Henriques, R.; Ribeiro, C.F.; Seica, R.; Cruz, T.; Cabral, C.; Fernandes, R.; Piedade, F.; Robalo, M.P.; et al. Improvement of Glycaemia and Endothelial Function by a New Low-Dose Curcuminoid in an Animal Model of Type 2 Diabetes. Int. J. Mol. Sci. 2022, 23, 5652. [Google Scholar] [CrossRef] [PubMed]
  217. Den Hartogh, D.J.; Vlavcheski, F.; Giacca, A.; Tsiani, E. Attenuation of Free Fatty Acid (FFA)-Induced Skeletal Muscle Cell Insulin Resistance by Resveratrol is Linked to Activation of AMPK and Inhibition of mTOR and p70 S6K. Int. J. Mol. Sci. 2020, 21, 4900. [Google Scholar] [CrossRef] [PubMed]
  218. Vlavcheski, F.; Den Hartogh, D.J.; Giacca, A.; Tsiani, E. Amelioration of High-Insulin-Induced Skeletal Muscle Cell Insulin Resistance by Resveratrol Is Linked to Activation of AMPK and Restoration of GLUT4 Translocation. Nutrients 2020, 12, 914. [Google Scholar] [CrossRef]
  219. Zhang, Y.J.; Zhao, H.; Dong, L.; Zhen, Y.F.; Xing, H.Y.; Ma, H.J.; Song, G.Y. Resveratrol ameliorates high-fat diet-induced insulin resistance and fatty acid oxidation via ATM-AMPK axis in skeletal muscle. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 9117–9125. [Google Scholar] [CrossRef]
  220. Szkudelska, K.; Deniziak, M.; Hertig, I.; Wojciechowicz, T.; Tyczewska, M.; Jaroszewska, M.; Szkudelski, T. Effects of Resveratrol in Goto-Kakizaki Rat, a Model of Type 2 Diabetes. Nutrients 2019, 11, 2488. [Google Scholar] [CrossRef]
  221. Seyyedebrahimi, S.; Khodabandehloo, H.; Nasli Esfahani, E.; Meshkani, R. The effects of resveratrol on markers of oxidative stress in patients with type 2 diabetes: A randomized, double-blind, placebo-controlled clinical trial. Acta Diabetol. 2018, 55, 341–353. [Google Scholar] [CrossRef]
  222. Delpino, F.M.; Figueiredo, L.M. Resveratrol supplementation and type 2 diabetes: A systematic review and meta-analysis. Crit. Rev. Food Sci. Nutr. 2022, 62, 4465–4480. [Google Scholar] [CrossRef]
  223. Jeyaraman, M.M.; Al-Yousif, N.S.H.; Singh Mann, A.; Dolinsky, V.W.; Rabbani, R.; Zarychanski, R.; Abou-Setta, A.M. Resveratrol for adults with type 2 diabetes mellitus. Cochrane Database Syst. Rev. 2020, 1, CD011919. [Google Scholar] [CrossRef]
  224. Zeraattalab-Motlagh, S.; Jayedi, A.; Shab-Bidar, S. The effects of resveratrol supplementation in patients with type 2 diabetes, metabolic syndrome, and nonalcoholic fatty liver disease: An umbrella review of meta-analyses of randomized controlled trials. Am. J. Clin. Nutr. 2021, 114, 1675–1685. [Google Scholar] [CrossRef]
  225. Mason, S.A.; Parker, L.; van der Pligt, P.; Wadley, G.D. Vitamin C supplementation for diabetes management: A comprehensive narrative review. Free Radic. Biol. Med. 2023, 194, 255–283. [Google Scholar] [CrossRef]
  226. Boonthongkaew, C.; Tong-Un, T.; Kanpetta, Y.; Chaungchot, N.; Leelayuwat, C.; Leelayuwat, N. Vitamin C supplementation improves blood pressure and oxidative stress after acute exercise in patients with poorly controlled type 2 diabetes mellitus: A randomized, placebo-controlled, cross-over study. Chin. J. Physiol. 2021, 64, 16–23. [Google Scholar] [CrossRef] [PubMed]
  227. Liu, M.; Park, S. A Causal Relationship between Vitamin C Intake with Hyperglycemia and Metabolic Syndrome Risk: A Two-Sample Mendelian Randomization Study. Antioxidants 2022, 11, 857. [Google Scholar] [CrossRef] [PubMed]
  228. Mason, S.A.; Keske, M.A.; Wadley, G.D. Effects of Vitamin C Supplementation on Glycemic Control and Cardiovascular Risk Factors in People with Type 2 Diabetes: A GRADE-Assessed Systematic Review and Meta-Analysis of Randomized Controlled Trials. Diabetes Care 2021, 44, 618–630. [Google Scholar] [CrossRef] [PubMed]
  229. Benetti, E.; Mastrocola, R.; Chiazza, F.; Nigro, D.; D’Antona, G.; Bordano, V.; Fantozzi, R.; Aragno, M.; Collino, M.; Minetto, M.A. Effects of vitamin D on insulin resistance and myosteatosis in diet-induced obese mice. PLoS ONE 2018, 13, e0189707. [Google Scholar] [CrossRef] [PubMed]
  230. Chang, E. 1,25-Dihydroxyvitamin D Decreases Tertiary Butyl-Hydrogen Peroxide-Induced Oxidative Stress and Increases AMPK/SIRT1 Activation in C2C12 Muscle Cells. Molecules 2019, 24, 3903. [Google Scholar] [CrossRef] [PubMed]
  231. Farahmand, M.A.; Daneshzad, E.; Fung, T.T.; Zahidi, F.; Muhammadi, M.; Bellissimo, N.; Azadbakht, L. What is the impact of vitamin D supplementation on glycemic control in people with type-2 diabetes: A systematic review and meta-analysis of randomized controlled trails. BMC Endocr. Disord. 2023, 23, 15. [Google Scholar] [CrossRef] [PubMed]
  232. Hands, J.M.; Patrick, R.; Frame, L.A. Vitamin D and Risk for Type 2 Diabetes in People with Prediabetes. Ann. Intern. Med. 2023, 176, eL230201. [Google Scholar] [CrossRef] [PubMed]
  233. Cojic, M.; Kocic, R.; Klisic, A.; Kocic, G. The Effects of Vitamin D Supplementation on Metabolic and Oxidative Stress Markers in Patients with Type 2 Diabetes: A 6-Month Follow Up Randomized Controlled Study. Front. Endocrinol. 2021, 12, 610893. [Google Scholar] [CrossRef] [PubMed]
  234. Gu, J.C.; Wu, Y.G.; Huang, W.G.; Fan, X.J.; Chen, X.H.; Zhou, B.; Lin, Z.J.; Feng, X.L. Effect of vitamin D on oxidative stress and serum inflammatory factors in the patients with type 2 diabetes. J. Clin. Lab. Anal. 2022, 36, e24430. [Google Scholar] [CrossRef] [PubMed]
  235. Kazemi, A.; Ryul Shim, S.; Jamali, N.; Hassanzadeh-Rostami, Z.; Soltani, S.; Sasani, N.; Mohsenpour, M.A.; Firoozi, D.; Basirat, R.; Hosseini, R.; et al. Comparison of nutritional supplements for glycemic control in type 2 diabetes: A systematic review and network meta-analysis of randomized trials. Diabetes Res. Clin. Pract. 2022, 191, 110037. [Google Scholar] [CrossRef] [PubMed]
  236. Whytock, K.L.; Pino, M.F.; Sun, Y.; Yu, G.; De Carvalho, F.G.; Yeo, R.X.; Vega, R.B.; Parmar, G.; Divoux, A.; Kapoor, N.; et al. Comprehensive interrogation of human skeletal muscle reveals a dissociation between insulin resistance and mitochondrial capacity. Am. J. Physiol. Endocrinol. Metab. 2023, 325, E291–E302. [Google Scholar] [CrossRef]
  237. Wikstrom, J.D.; Katzman, S.M.; Mohamed, H.; Twig, G.; Graf, S.A.; Heart, E.; Molina, A.J.; Corkey, B.E.; de Vargas, L.M.; Danial, N.N.; et al. beta-Cell mitochondria exhibit membrane potential heterogeneity that can be altered by stimulatory or toxic fuel levels. Diabetes 2007, 56, 2569–2578. [Google Scholar] [CrossRef] [PubMed]
  238. Herman, W.H.; Ratner, R.E. Metformin Should Be Used to Treat Prediabetes in Selected Individuals. Diabetes Care 2020, 43, 1988–1990. [Google Scholar] [CrossRef]
  239. de Marañón, A.M.; Canet, F.; Abad-Jiménez, Z.; Jover, A.; Morillas, C.; Rocha, M.; Victor, V.M. Does Metformin Modulate Mitochondrial Dynamics and Function in Type 2 Diabetic Patients? Antioxid. Redox Signal. 2021, 35, 377–385. [Google Scholar] [CrossRef]
  240. de Marañón, A.M.; Díaz-Pozo, P.; Canet, F.; Díaz-Morales, N.; Abad-Jiménez, Z.; López-Domènech, S.; Vezza, T.; Apostolova, N.; Morillas, C.; Rocha, M.; et al. Metformin modulates mitochondrial function and mitophagy in peripheral blood mononuclear cells from type 2 diabetic patients. Redox Biol. 2022, 53, 102342. [Google Scholar] [CrossRef]
  241. Li, A.; Zhang, S.; Li, J.; Liu, K.; Huang, F.; Liu, B. Metformin and resveratrol inhibit Drp1-mediated mitochondrial fission and prevent ER stress-associated NLRP3 inflammasome activation in the adipose tissue of diabetic mice. Mol. Cell Endocrinol. 2016, 434, 36–47. [Google Scholar] [CrossRef]
  242. Marshall, S.M. 60 years of metformin use: A glance at the past and a look to the future. Diabetologia 2017, 60, 1561–1565. [Google Scholar] [CrossRef]
Figure 1. Phenotypic and Clinical manifestation of T2D heterogenous populations. Classification of T2D based upon specific characteristics features including fat accumulation, insulin action, β-cell function, multi-omics biomarkers and disease progression. Created with BioRender.com.
Figure 1. Phenotypic and Clinical manifestation of T2D heterogenous populations. Classification of T2D based upon specific characteristics features including fat accumulation, insulin action, β-cell function, multi-omics biomarkers and disease progression. Created with BioRender.com.
Ijms 25 01504 g001
Figure 2. Mechanisms of chronic ROS/RNS generation in dysfunctional mitochondria. Exposure of cells to diabetogenic conditions like hypoxia, inflammation, hyperglycemia, and overproduction of free fatty acids can lead to increased leakage of electrons from the mitochondrial ETC. This enhanced electron leakage results in the overproduction of ROS/RNS, which can lead to oxidative stress and cellular damage. Created with BioRender.com.
Figure 2. Mechanisms of chronic ROS/RNS generation in dysfunctional mitochondria. Exposure of cells to diabetogenic conditions like hypoxia, inflammation, hyperglycemia, and overproduction of free fatty acids can lead to increased leakage of electrons from the mitochondrial ETC. This enhanced electron leakage results in the overproduction of ROS/RNS, which can lead to oxidative stress and cellular damage. Created with BioRender.com.
Ijms 25 01504 g002
Figure 3. Transient ROS signaling in the healthy islet β-cell. (1) Glucose metabolism increases the levels of ROS in a transient manner, which triggers signaling events that ultimately culminating in GSIS. (2) The increased levels of ROS (H2O2) activate ryanodine receptors (RyRs) at the surface of endoplasmic reticulum, which causes the release of Ca2+ sufficient enough to trigger GSIS. (3) NOX4 at the mitochondrial surface causes an increase in ROS that causes the closure of ATP sensitive potassium channel and ensuing GSIS. (4) The plasma membrane associated NOX2 activation requires GTPase signaling cascades that lead to mobilization of insulin secretory granules towards the plasma membrane for release of insulin. Created with BioRender.com.
Figure 3. Transient ROS signaling in the healthy islet β-cell. (1) Glucose metabolism increases the levels of ROS in a transient manner, which triggers signaling events that ultimately culminating in GSIS. (2) The increased levels of ROS (H2O2) activate ryanodine receptors (RyRs) at the surface of endoplasmic reticulum, which causes the release of Ca2+ sufficient enough to trigger GSIS. (3) NOX4 at the mitochondrial surface causes an increase in ROS that causes the closure of ATP sensitive potassium channel and ensuing GSIS. (4) The plasma membrane associated NOX2 activation requires GTPase signaling cascades that lead to mobilization of insulin secretory granules towards the plasma membrane for release of insulin. Created with BioRender.com.
Ijms 25 01504 g003
Figure 4. Differential mechanisms involved in the contribution of mitophagy associated with various stages of skeletal muscle IR. Mitophagy is a specialized form of macro-autophagy that selectively disposes of aging, damaged or dysfunctional mitochondria via autophagolysosomes-mediated degradation. (a) Ubiquitin-dependent mitophagy relies upon the stabilization of PINK1 on the OMM upon mitochondrial membrane damage and/or decrease of membrane potential. PINK1 autoactivation via Ser 228 and 405 autophosphorylation enables Ser 65 phosphorylation of ubiquitin on OMM proteins to recruit PARKIN. PINK1 activation of PARKIN via Ser 65 phosphorylation activates its E3-ubiquitin ligase activity triggering a feedback loop of further Ser 65 phosphorylation of polyubiquitin chains on OMM proteins and recruitment of more PARKIN. The ser-65 phosphorylated polyubiquitin chains of the OMM proteins facilitate LC3-positive autophagosome recruitment via LC3 interacting adaptor proteins. (b) Receptor-dependent mitophagy relies upon OMM and inner mitochondrial membrane (IMM) proteins with LC3-interacting motif regions such as FUNDC1 and BNIP3L/NIX that directly bind to LC3 enabling mitochondria-autophagosome fusion. (c) Proposed model for mitophagy role in T2D where increased mitophagy acts as a protective early response to protect against accumulative mitochondrial damage via ROS production, whilst T2D muscle the decreased mitophagy activity enables damaged mitochondria accumulation. Whether STX4 regulates mitophagy via a PARKIN-dependent or -independent pathway remains in question (?). Created with BioRender.com.
Figure 4. Differential mechanisms involved in the contribution of mitophagy associated with various stages of skeletal muscle IR. Mitophagy is a specialized form of macro-autophagy that selectively disposes of aging, damaged or dysfunctional mitochondria via autophagolysosomes-mediated degradation. (a) Ubiquitin-dependent mitophagy relies upon the stabilization of PINK1 on the OMM upon mitochondrial membrane damage and/or decrease of membrane potential. PINK1 autoactivation via Ser 228 and 405 autophosphorylation enables Ser 65 phosphorylation of ubiquitin on OMM proteins to recruit PARKIN. PINK1 activation of PARKIN via Ser 65 phosphorylation activates its E3-ubiquitin ligase activity triggering a feedback loop of further Ser 65 phosphorylation of polyubiquitin chains on OMM proteins and recruitment of more PARKIN. The ser-65 phosphorylated polyubiquitin chains of the OMM proteins facilitate LC3-positive autophagosome recruitment via LC3 interacting adaptor proteins. (b) Receptor-dependent mitophagy relies upon OMM and inner mitochondrial membrane (IMM) proteins with LC3-interacting motif regions such as FUNDC1 and BNIP3L/NIX that directly bind to LC3 enabling mitochondria-autophagosome fusion. (c) Proposed model for mitophagy role in T2D where increased mitophagy acts as a protective early response to protect against accumulative mitochondrial damage via ROS production, whilst T2D muscle the decreased mitophagy activity enables damaged mitochondria accumulation. Whether STX4 regulates mitophagy via a PARKIN-dependent or -independent pathway remains in question (?). Created with BioRender.com.
Ijms 25 01504 g004
Figure 5. Mechanisms of skeletal muscle ROS generation in T2D. Hyperglycemia-mediated mitochondrial defects and the activation of stress pathways increase ROS in skeletal muscle, which then contributes to islet cell damage/dysfunction. GSH: glutathione; SOD: superoxide dismutase; NADPH: nicotinamide adenine dinucleotide phosphate; AGE: advanced glycation end products; PKC: protein kinase C; DAG: diacyl glycerol; NF-κB: nuclear factor-kappa-B; PKC; NOX: NADPH oxidases; UDPGlcNAc: uridine diphosphate-N-acetylglucosamine. The (↑) indicates upregulation/increase, and (↓) arrow indicates downregulation/decrease in the stress associated pathways, signaling molecules and antioxidants. All the connecting arrows specify the sequence of cellular responses induced by ROS via elevated levels of extra/intracellular glucose and lipids. Created with BioRender.com.
Figure 5. Mechanisms of skeletal muscle ROS generation in T2D. Hyperglycemia-mediated mitochondrial defects and the activation of stress pathways increase ROS in skeletal muscle, which then contributes to islet cell damage/dysfunction. GSH: glutathione; SOD: superoxide dismutase; NADPH: nicotinamide adenine dinucleotide phosphate; AGE: advanced glycation end products; PKC: protein kinase C; DAG: diacyl glycerol; NF-κB: nuclear factor-kappa-B; PKC; NOX: NADPH oxidases; UDPGlcNAc: uridine diphosphate-N-acetylglucosamine. The (↑) indicates upregulation/increase, and (↓) arrow indicates downregulation/decrease in the stress associated pathways, signaling molecules and antioxidants. All the connecting arrows specify the sequence of cellular responses induced by ROS via elevated levels of extra/intracellular glucose and lipids. Created with BioRender.com.
Ijms 25 01504 g005
Figure 6. Islet β-cell-skeletal muscle crosstalk is mediated by EVs. During exercise (a) or IR (b), skeletal muscles secrete EVs which contain numerous secreted factors, including mitochondrial cargo. These factors can impact the health of islet β-cells. Refer to Table 2 for more details. GLP-1, Glucagon-like peptide-1. Solid arrows indicate the release of EVs from the tissues (skeletal muscle and islet β-cell). The dashed arrow and (?) indicates that it remains to be determined as to whether EVs released from an islet β-cell are required to regulate skeletal muscle function. Created with BioRender.com.
Figure 6. Islet β-cell-skeletal muscle crosstalk is mediated by EVs. During exercise (a) or IR (b), skeletal muscles secrete EVs which contain numerous secreted factors, including mitochondrial cargo. These factors can impact the health of islet β-cells. Refer to Table 2 for more details. GLP-1, Glucagon-like peptide-1. Solid arrows indicate the release of EVs from the tissues (skeletal muscle and islet β-cell). The dashed arrow and (?) indicates that it remains to be determined as to whether EVs released from an islet β-cell are required to regulate skeletal muscle function. Created with BioRender.com.
Ijms 25 01504 g006
Figure 7. Adipose tissue Mito-EV mediated inter-organ communications. Mito-EV cargo impacts whole-body metabolism by influencing pancreatic β-cells and macrophage function. Negative effects (↓) on macrophage lead to impaired insulin stimulated glucose uptake by skeletal muscle myocytes and decreased GSIS from pancreatic β-cells. Positive effect (↑), increased insulin tolerance and sensitivity, or increased GSIS from pancreatic β-cells. Created with BioRender.com.
Figure 7. Adipose tissue Mito-EV mediated inter-organ communications. Mito-EV cargo impacts whole-body metabolism by influencing pancreatic β-cells and macrophage function. Negative effects (↓) on macrophage lead to impaired insulin stimulated glucose uptake by skeletal muscle myocytes and decreased GSIS from pancreatic β-cells. Positive effect (↑), increased insulin tolerance and sensitivity, or increased GSIS from pancreatic β-cells. Created with BioRender.com.
Ijms 25 01504 g007
Table 1. Skeletal muscle-secreted factors, linked to oxidative stress (or ROS), insulin resistance, or Type 2 diabetic conditions and their effects on pancreatic β-cells.
Table 1. Skeletal muscle-secreted factors, linked to oxidative stress (or ROS), insulin resistance, or Type 2 diabetic conditions and their effects on pancreatic β-cells.
Myokines or miRNA from Skeletal Muscle EVsRelevance with Oxidative Stress (ROS), Insulin Resistance (IR) or Type 2 Diabetes (T2D)Effect on Pancreatic β-CellsRef.
FollistatinReduces ROS in T2DEnhances GSIS and β-cell proliferation[194]
IrisinReduces inflammation-mediated oxidative stress Enhances insulin production, GSIS and β-cell proliferation[195]
CX3CL1Reduces ROSEnhances GSIS[196,197]
miR-133aExpression increased during oxidative stressReduces GSIS[198]
CXCL10Enhanced by oxidative stress or IREnhances apoptosis and reduces β-cell proliferation[199]
Table 2. Mitochondrial cargo within Mito-EVs linked to oxidative stress, insulin resistance or Type 2 diabetic conditions and their effects on recipient cells.
Table 2. Mitochondrial cargo within Mito-EVs linked to oxidative stress, insulin resistance or Type 2 diabetic conditions and their effects on recipient cells.
Cell SourceMitochondrial Cargo within Mito-EVsRelevance to ROS, IR, and T2D Effects on Recipient CellsRef.
β-cellsPHGDH, ALDH9A1, ADSL, and DNML1Increase inflammation-mediated oxidative stressHorizontal transfer of EVs promoted immune cell recruitment to islets insulin-positive β-cells[200]
Adipocytes Fatty acid oxidation enzymes ECHA and HCDHTumorigenicity exacerbated in obesityAdipocyte-EVs promoted melanoma tumor cell progression via transfer of fatty acids under obesity conditions in vivo and ex vivo[201]
AdipocytesHSP60, VDAC1, COXIVPromote ROS under obesity Induce mitochondrial dysfunction of the host network and free radical production in cardiomyocytes. A beneficial effect that protects the heart from ischemia reperfusion injury [202]
Adipocytes mtDNA, AMP, ATP, that are MitoTracker Green (MTG) positive, without changes to PDHβIncreased ROS Adapt EVs as a quality control to remove damaged mitochondria [203]
Adipose tissue MSCs and adipocytes PGM1, PCX, MDH1Insulin secretion under obesity-induced IR/T2DPromote GSIS in β-cells [204]
Adipocytes HSP70 and HSP90Promote IR in myocytesInfluence monocyte differentiation into pro-inflammatory macrophages, which directly act on myocytes to impair insulin-stimulated glucose uptake and insulin signaling in vitro and in vivo[205]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Veluthakal, R.; Esparza, D.; Hoolachan, J.M.; Balakrishnan, R.; Ahn, M.; Oh, E.; Jayasena, C.S.; Thurmond, D.C. Mitochondrial Dysfunction, Oxidative Stress, and Inter-Organ Miscommunications in T2D Progression. Int. J. Mol. Sci. 2024, 25, 1504. https://doi.org/10.3390/ijms25031504

AMA Style

Veluthakal R, Esparza D, Hoolachan JM, Balakrishnan R, Ahn M, Oh E, Jayasena CS, Thurmond DC. Mitochondrial Dysfunction, Oxidative Stress, and Inter-Organ Miscommunications in T2D Progression. International Journal of Molecular Sciences. 2024; 25(3):1504. https://doi.org/10.3390/ijms25031504

Chicago/Turabian Style

Veluthakal, Rajakrishnan, Diana Esparza, Joseph M. Hoolachan, Rekha Balakrishnan, Miwon Ahn, Eunjin Oh, Chathurani S. Jayasena, and Debbie C. Thurmond. 2024. "Mitochondrial Dysfunction, Oxidative Stress, and Inter-Organ Miscommunications in T2D Progression" International Journal of Molecular Sciences 25, no. 3: 1504. https://doi.org/10.3390/ijms25031504

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