Next Article in Journal
Single-Cell RNA-Sequencing: Opening New Horizons for Breast Cancer Research
Previous Article in Journal
Identification of Shaker Potassium Channel Family Members and Functional Characterization of SsKAT1.1 in Stenotaphrum secundatum Suggest That SsKAT1.1 Contributes to Cold Resistance
Previous Article in Special Issue
Acute Hyperglycemia-Induced Injury in Myocardial Infarction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diabetic Cardiomyopathy: Role of Cell Death, Exosomes, Fibrosis and Epicardial Adipose Tissue

by
Antonella Galeone
1,
Alessia Annicchiarico
2,
Cinzia Buccoliero
2,
Barbara Barile
2,
Giovanni Battista Luciani
1,
Francesco Onorati
1,
Grazia Paola Nicchia
2 and
Giacomina Brunetti
2,*
1
Department of Surgery, Dentistry, Pediatrics and Gynecology, Division of Cardiac Surgery, University of Verona, 37129 Verona, Italy
2
Department of Biosciences, Biotechnologies and Environment, University of Bari Aldo Moro, 70125 Bari, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(17), 9481; https://doi.org/10.3390/ijms25179481 (registering DOI)
Submission received: 31 July 2024 / Revised: 21 August 2024 / Accepted: 29 August 2024 / Published: 31 August 2024
(This article belongs to the Special Issue Advances in Molecular Research of Diabetic Cardiomyopathy)

Abstract

:
Diabetic cardiomyopathy (DCM) represents one of the typical complications associated with diabetes. It has been described as anomalies in heart function and structure, with consequent high morbidity and mortality. DCM development can be described by two stages; the first is characterized by left ventricular hypertrophy and diastolic dysfunction, and the second by heart failure (HF) with systolic dysfunction. The proposed mechanisms involve cardiac inflammation, advanced glycation end products (AGEs) and angiotensin II. Furthermore, different studies have focused their attention on cardiomyocyte death through the different mechanisms of programmed cell death, such as apoptosis, autophagy, necrosis, pyroptosis and ferroptosis. Exosome release, adipose epicardial tissue and aquaporins affect DCM development. This review will focus on the description of the mechanisms involved in DCM progression and development.

1. Introduction

Diabetic cardiomyopathy (DCM) represents one of the typical complications associated with diabetes. It has been described as anomalies in heart function and structure, with consequent high morbidity and mortality. However, the pathophysiological mechanisms of DCM can be different according to the type of diabetes. It is important to remember that the two main types of diabetes are type 1 diabetes (T1D) and type 2 diabetes (T2D). The first is characterized by insulin deficiency [1], whereas the latter is characterized by insulin resistance [2]. Consistently, DCM is associated with left ventricular mass increase [3] that in T1D may be due to the younger age of disease development, whereas in T2D, DCM is independent of factors such as race and obesity [4]. In the attempt to deepen the mechanisms, researchers focus on insulin signaling, which involves two interacting pathways: the MAPK pathway and the insulin receptor substrate 1 (IRS-1) pathway [5]. Obviously, in T2D, insulin resistance alters the balance, shifting it in favor of the MAPK pathway with consequent effects on cell metabolism and growth leading to cardiac fibrosis and diastolic dysfunction.
DCM development can be described by two stages; the first is characterized by left ventricular hypertrophy and diastolic dysfunction, and the second by heart failure (HF) with systolic dysfunction [6]. The proposed mechanisms involve cardiac inflammation, oxidative stress, advanced glycation end products (AGEs) and angiotensin II (Ang-II) [7,8]. Furthermore, different studies have focused their attention on cardiomyocyte death through the different mechanisms of programmed cell death, such as apoptosis, autophagy, necrosis, pyroptosis and ferroptosis [9,10]. In this review, we will detail the mechanisms associated with these processes (Figure 1).

2. Cardiomyocyte Cell Death

2.1. Cardiomyocyte Apoptosis

Cardiomyocyte apoptosis may occur in the initial stage of DCM development and is linked to myocardial hypertrophy and HF. The mechanisms reported endoplasmic reticulum (ER) stress as a crucial event for cardiomyocyte apoptosis in DCM. Positive effects have been reported for chlorogenic acid and hydrogen sulfide to counteract ER stress using in vitro and in vivo models. Oxidative stress in myocardial ER can be activated by permanent hyperglycemia, resulting in caspase activation [11,12,13]. Consistently, high glucose levels have been linked to caspase-8 and -9 in cardiomyocytes in neonatal rats [14].
An additional factor affecting cardiomyocyte apoptosis in DCM is represented by chronic inflammation, which leads to the increased production of nitric oxide (NO) and sustains the expression of proto-oncogenes with the consequent development of ventricular diastolic dysfunction [15]. The altered levels of glucose in DCM patients’ blood determine the activation of leukocytes together with the recruitment of activated monocytes, neutrophils and macrophages. Consequently, the levels of pro-inflammatory cytokines are also increased. Consistently, high levels of Interleukin (IL)-1β, IL-6, Tumor Necrosis Factor (TNF)-α and Tumor Growth Factor (TGF)-β1 lead to cardiomyocyte apoptosis and thus DCM [16]. Inflammation can be also exacerbated by increased levels of ROS and NLR family pyrin domain containing 3 (NLRP3) inflammasome. Consistently, the dysregulation of the latter is associated with an immunomodulatory response in DCM; in contrast, NLRP3 silencing improves cardiac remodeling in diabetes and alters cardiac function [17].
A key role has also been attributed to intracellular Ca2+ levels, which are important for cardiomyocyte contraction. Interestingly, the abnormal gene transduction of Ca2+-ATPase has been demonstrated in diabetic rats, which displayed a reduced uptake of Ca2+ in the sarcoplasmic reticulum together with a reduction in the Na+-Ca2+ transporter in the cardiomyocyte membrane, leading to an increase in Ca2+ in the cells, with a consequent increased duration of the action potential and a shortened systolic phase, together with a prolonged diastolic phase [18,19,20]. Ca2+ overload is also associated with an increased uptake by mitochondria, which is linked to cardiomyocyte apoptosis [21].

2.2. Cardiomyocytes Pyroptosis

Pyroptosis and apoptosis represent two distinct types of programmed cell death with some similarities [22]. In detail, pyroptosis is associated with membrane destruction, cell swelling, intact nucleus, inflammation, pore formation, pyroptotic body formation and the activation of caspase-1, caspase-4, -5, -11 and -12, whereas apoptosis is characterized by apoptotic body formation and caspase-2, -7 and -10 activation [23,24,25]. Pyroptosis can sustain DCM development through the involvement of different pathways, and myocardial cells can die by pyroptosis to speed up DCM development. Different pathways can be involved in DCM pyroptosis, such as NF-κB/NLRP3, ROS and Nuclear factor erythroid 2-related factor 2 (Nrf-2) [26,27,28,29,30]. Hyperglycemia leads to elevated ROS production, promoting the binding of TXN1P to NLRP3, and consequently the activation of the inflammasome, which is also a crucial mediator of DCM progression [31].

2.3. Autophagy

The word autophagy arises from Greek and characterizes the self-degradation and recycling of cellular components [32]. In heart homeostasis, autophagy represents a dynamic process leading to eliminated misfolded proteins, damaged organelles and cellular debris [33,34]. It is a tightly regulated process fundamental for cellular adaptation, survival and renewal. Consistently, hyperglycemia impairs autophagy by inhibiting autophagosome formation and maturation, with the consequent accumulation of unuseful material in cardiomyocytes, thus impairing cardiac activity in diabetes [35,36]. Oxidative stress also impairs autophagy through the disruption of redox-sensitive pathways. An additional modulator of autophagy is represented by inflammation, which contributes to the alteration of cardiac remodeling dysfunction in DCM [37,38].
A key role in autophagy has been demonstrated by noncoding RNAs, including microRNAs (miRNAs), long noncoding RNAs (lncRNAs) and circular RNAs. LncRNAs have a length of about 200 nucleotides, lack protein-coding properties, and regulate gene expression and different cellular processes [32,39]. Among these, the lncRNA Diabetic Cardiomyopathy-Related Factor (DCRF) is involved in DCM pathogenesis [40]. This factor affects gene expression and cellular activities in the cardiomyocytes of diabetic patients. Consistently, using a rat model of DCM, it has been demonstrated that DCRF knockdown led to diminished cardiomyocyte autophagy, ameliorated cardiac function and reduced myocardial fibrosis. Increased DCRF levels were linked to hyperglycemia [32]. An additional important lncRNA is represented by H19. Using a rat model, the role of H19 has been evaluated in DCM. In diabetic rats, overexpression led to reduced cardiomyocyte autophagy and increased left ventricular performance [41,42].
GAS5 is involved in the regulation of cellular growth and apoptosis. In DCM, GAS5 attenuated myocardial damage by stimulating cardiomyocyte autophagy through the modulation of the miR221-3p/p27 axis [43,44]. NEAT1 is a lncRNA localized in the nucleus that participates in nuclear body formation. It is involved in numerous cellular processes, such as nuclear structure maintenance and gene expression modulation. In diabetic rats, Neat1 overexpression determined increased serum myocardial enzyme levels, reduced superoxide dismutase concentration and cardiomyocyte viability, and augmented infarct size. In diabetic rats, enhanced Neat1 levels affected cardiac ischemia/reperfusion damage through the induction of autophagy [45,46].
miRNAs have been implicated in different cellular processes, both in health and disease. Their dysregulation has been implicated in different diseases including DCM. The role of miR-200a-3p in DCM has been investigated using a model of DCM realized with db/db mice. These mice displayed reduced levels of miR-200a-3p in the heart. Diabetic mice displayed the overexpression of this miRNA, with consequent improved autophagy and decreased myocardial injury, cardiac dysfunction, apoptosis, fibrosis and inflammation. In this model, the increased levels of miR-200a-3p are linked to DCM in T2D by modulating the Mst1/Sirt3/AMPK axis acting on the expression of FOXO3 [47].
Another important miRNA is represented by miR-207; its role in T2D has been studied in DCM animal models. They displayed autophagy malfunction, enhanced cell apoptosis and affected miR-207 levels. In cardiomyocytes, miR-207 inhibited autophagy and enhanced apoptosis. Additionally, in cardiomyocytes, miR-207 targeted LAMP2, an important autophagy-related protein, to inhibit autophagy and promote apoptosis, thus promoting DCM in T2D [48,49].
The role of miR-30d has been evaluated in DCM, showing that it inhibited autophagy in rats by modulating the KLF9/VEGFA pathway [50].
Other authors reported that miRNA expression can be affected by glycemic management but cannot be reversed to control the condition [51].

2.4. Ferroptosis

Ferroptosis is an iron-dependent nonapoptotic cell death associated with lipid-peroxidation accumulation, leading to the release of extremely cytotoxic products such as malondialdehyde and 4-hydroxynonenal, which damage cell membranes, proteins and DNA [52]. Different mediators have been demonstrated to be involved, such as CD36, Nrf2 and Zinc finger antisense (ZFAS1). In detail, CD36 is a multifunctional receptor involved in lipid metabolism and transport, energy metabolism reprogramming and inflammatory response. CD36 expression increases in DCM cardiomyocytes.
The transcription factor Nrf2 modulates the transcription of different factors, such as scavenger receptors, antioxidant genes, and autophagic and transporter proteins [53]. In vivo, it works as an antioxidant stress regulator. In DCM, Nrf2 is associated with ferroptosis in cardiomyocytes [54]. It has been demonstrated that DCM could activate Nrf2 by blocking autophagy, as well as Nrf2-mediated iron overload and lipid peroxidation, that in turn activated ferroptosis in cardiomyocytes [55].
ZFAS1 is a new lncRNA involved in numerous diseases. Its upregulation, together with miR-150-5p levels and ferroptosis, can be found in high glucose-treated cardiomyocytes and DCM animal models [56]. Interestingly, Cyclin D2 can modulate miR-150-5p. Cyclin 2 overexpression inhibited ferroptosis, whereas its downregulation stimulated ferroptosis [56].

3. Aquaporins and DCM

Aquaporins (AQPs) are a ubiquitous family of water channels that mediate the transport of water and other small molecules across cell membranes [57]. The major AQPs in the human cardiovascular system are AQP1, expressed in arteries, endothelia, myocytes and vascular smooth muscle cells [58]; AQP3, AQP4 and AQP7, localized in cardiomyocytes; and AQP9, found in endothelial cells.
Despite the widely recognized role of AQPs in cardiovascular physiology and pathology [58,59,60], very little is known about their role in DCM. AQP1 has been identified as an emerging protein involved in the pathophysiology of myocardial edema, coronary atherosclerosis [58] and also DCM, where its expression was found to be detrimentally reduced in high glucose-evoked cardiomyocyte injury [61]. It has been proven that the inhibition of miR-1306–5p upregulates AQP1 and ameliorates the derived injury. Similarly, AQP1 and AQP4 are likely to play a beneficial role in DCM resulting from a pharmacological study evaluating the effects of the antidiabetic treatment Empaglifozin (EMPA) in type 2 diabetic rats [61]. In this study, the administration of EMPA was found to be associated with beneficial cardioprotective effects, such as decreased fibrosis, apoptosis and edema, likely due to the upregulation of the water channels and, therefore, fine water balance control [61]. To our knowledge, these are the only studies that provide evidence of the involvement of water channels in diabetic cardiomyopathy. Further research is needed to better understand their contribution to this promising field of study.

4. DCM and Exosomes

Exosomes, small extracellular vesicles of 30–100 nm in size, have been studied for several years in different areas of research from tumors to cardiovascular diseases, prospecting their use in clinical applications [62,63,64,65,66,67,68]. Exosomes can carry proteins, RNAs (mRNA, miRNA and noncoding RNA) and DNA sequences of great interest. Exosomes are studied for their ability to act in intercellular communication. For example, noncoding RNAs released from tumor-derived exosomes induce the polarization of M1 (pro-inflammatory) macrophages to M2 (anti-inflammatory) phenotypes, enhancing the suppression of immune cells [69]. Of great interest is the therapeutic potential of exosomes in different diseases including T2D, cutaneous wound healing, kidney, ocular and Alzheimer’s disease [70,71,72,73,74,75,76,77]; in cancer therapy resistance to chemotherapy agents [78]; in drug delivery [79,80,81]; and in autoimmune diseases including rheumatoid arthritis (RA), systemic lupus erythematosus (SLE), T1D, Sjogren’s syndrome (SS), multiple sclerosis (MS), inflammatory bowel disease (IBD) and systemic sclerosis (SSc) [82]. Furthermore, exosomes promote rotator cuff tendon–bone healing [83] and cerebral ischemia repair [84]; improve the repair of diabetic ischemia of the hind limb [85] and tendinopathy [86]; and promote skin wound healing [70,87]. Moreover, the exosomes’ role in heart disease, including DCM, and related clinical studies and therapy has been investigated [88,89,90,91,92].
Chaturvedi et al. investigated the benefit of cardiosomes, exosomes released from cardiomyocytes, during physical exercise on the cardiovascular complications of diabetes [93]. The study focuses on the ability of exercise to reduce the levels of matrix metalloprotease 9 (MMP9) in db/db mice as models of T2D and on the underlying molecular mechanism. Notably, Chaturvedi hypothesized the release of specific cardiomyosome microRNAs (mir455, mir29b, mir323-5p and mir466), which, binding to the 3′ region of MMP9, downregulated its expression and mitigated extracellular matrix remodelling [93]. Moreover, Hirai et al. investigated the role of cardiosphere-derived exosomes in myocardial repair in pediatric cardiomyopathy [94]. Of note, cardiosphere-derived exosomes, enriched with miR-146a-5p, inhibit myocyte apoptosis and fibrosis, enhancing angiogenesis and cardiac activity after infarction [94]. Notably, Gan et al. studied the role of circulating extracellular vesicles isolated from the serum of mice fed either normal or high-fat diets in exacerbating myocardial ischemia/reperfusion injury [95]. Compared to the control, the intramyocardial injection of serum vesicles from the animals fed a high-fat diet significantly increased myocardial ischemia/reperfusion (MI/R) injury in mice [95]. This result was confirmed by the poor recovery of cardiac function, larger infarct size and increased death by apoptosis [95]. In contrast, the injection of vesicles from the animals fed a normal diet had an opposite effect in reducing myocardial ischemia/reperfusion injury [95]. Moreover, the intramyocardial injection of diabetic adipocyte vesicles and high glucose/high lipid-challenged nondiabetic adipocytes exacerbated MI/R damage. Notably, miR-130b-3p levels were significantly increased in all the above vesicles, and the intramyocardial administration of miR-130b-3p significantly increased MI/R injury in the nondiabetic mice, whereas miR-130b-3p inhibitors significantly attenuated MI/R injury in the diabetic mice [95]. Of great interest is the possibility of using stem cell-derived exosomes in treating cardiovascular diseases and in promoting cardiac repair, enhancing angiogenesis and reducing apoptosis [96].
Yu et al. reported that exosomes released from mesenchymal stem cells expressing high levels of GATA4 had cardioprotective capabilities, preserving cardiac contraction and reducing infarct size. In this regard, miR-19a was higher in cardiomyocytes and myocardia treated with exosomes derived from mesenchymal stem cells overexpressing GATA-4 than in those treated with exosomes derived from control mesenchymal stem cells [97].
In addition, it has been shown that exosomes, depending on the content released, can induce positive or negative effects on the myocardium [98,99]. In diabetic cardiomyopathy, high levels of miR-320 have been detected in cardiomyocyte-derived exosomes, adversely affecting the heart [100]. The use of an exosome secretion inhibitor, such as GW4869, could be a potential therapeutic strategy to mitigate exosome-mediated cardiac dysfunction in diabetic hearts [100,101,102]. Wang and colleagues demonstrated that exosomes released from diabetic cardiomyocytes contained detrimental substances, such as lower levels of Hsp20 than normal ones, implicated as a primary factor contributing to T1D- and T2D-induced organ damage including ventricular dysfunction, cardiac fibrosis and cardiomyocyte apoptotic death [101]. In addition to impaired cardiac function, exosomes released from diabetic cardiomyocytes can also mediate anti-angiogenesis events through the exosomal transfer of miR-320 into endothelial cells and embryonic development, since maternal exosomes in diabetes could cross the maternal–fetal barrier, promoting cardiac developmental deficits [102,103]. In this context, several studies have shown that cardiomyocyte-derived exosomes contain different mRNAs and miRNAs, proteins, and lipids, which can be released to adjacent cardiac endothelial cells, positively or negatively modulating their activity [102,104,105,106].
Exosomes derived from nondiabetic Wistar rat cardiomyocytes promoted the proliferation and migration of cardiac endothelial cells. In contrast, exosomes isolated from diabetic Goto–Kakizaki (GK) rat cardiomyocytes reduced the proliferation and migration of cardiac endothelial cells [107]. The contents of the exosomes isolated from diabetic GK cardiomyocytes and the exosomes from Wistar rats were investigated. Exosomes from diabetic GK cardiomyocytes had higher levels of miRNA-320 and lower levels of miRNA-126 and heat shock protein 20 (Hsp20) than exosomes isolated from nondiabetic Wistar rat cardiomyocytes [102]. miRNA-320 can be released to cardiac endothelial cells, downregulating the expression of IGF-1, Hsp20 and Ets-2 and negatively affecting the angiogenic role of the adjacent cardiac endothelial cells. In addition, Garcia and colleagues have shown that under conditions of glucose deprivation, immortalized H9C2 cardiomyocytes produce more exosomes whose cargo affects cardiac endothelial cell activity, inducing modification in the transcription of pro-angiogenic genes [108].

5. Cardiac Fibrosis

DCM is characterized by myocardial hypertrophy and fibrosis in the absence of coronary artery disease, hypertension, or valvular heart disease [109]. Cardiac fibrosis is a prominent feature of diabetic cardiomyopathy that increases myocardial stiffness and is associated with reduced diastolic function and systolic dysfunction that eventually leads to HF, and may promote arrhythmogenesis and a higher risk of sudden death [110]. Cardiac fibrosis is the consequence of the accumulation of the extracellular matrix (ECM) produced by cardiac fibroblasts [111]. Distinct types of myocardial fibrosis have been described, such as replacement fibrosis, interstitial fibrosis and perivascular fibrosis [111]. Replacement fibrosis is usually seen in myocardial infarction and refers to the formation of collagen-rich scar tissue in areas of myocardial necrosis [112]. Replacement fibrosis is the result of a reparative process following cardiomyocyte injury. Despite the absence of ischemic injury leading to myocardial necrosis in diabetic cardiomyopathy, abnormal metabolism represents a chronic myocardial cell injury and may induce cardiomyocyte apoptosis, leading to replacement fibrosis. Interstitial fibrosis refers to the deposition of the ECM in the endomysium and perimysium, while perivascular fibrosis indicates the expansion of periadventitial collagen in the cardiac microvasculature. In contrast to replacement fibrosis, interstitial and perivascular fibrosis are not related to cardiomyocyte death, but are rather the results of metabolic alterations, inflammation and oxidative stress. Cardiac fibroblasts are the main ECM-producing cells [113] and multiple pathways with a complex interplay are involved in the activation and proliferation of cardiac fibroblasts and the production of ECM proteins in diabetic cardiomyopathy, such as hyperglycemia, insulin resistance, AGEs, TGF-β, renin–angiotensin–aldosterone system (RAAS), and the imbalance between matrix metalloproteinases (MMPs) and tissue inhibitors of metalloproteinases (TIMPs) [109].
Hyperglycemia and insulin resistance promote the proliferation of cardiac fibroblasts and increase the production of ECM proteins. Experimental studies in vitro showed that a high-glucose environment stimulates cultured cardiac fibroblasts to synthesize large amounts of ECM proteins such as collagen and fibronectin [114]. Hyperglycemia also promotes the production of AGEs that bind to specific receptors on the cell membrane, induce the release of large amounts of reactive oxygen species (ROS) and activate nuclear factor-κ-gene binding (NF-кB), which is a transcription factor of various inflammatory factors such as TNF-α and IL-6 and regulates the expression of pro-fibrotic and hypertrophy-related genes [115]. AGEs also cause the cross-linking of myocardial collagen molecules to each other, leading to the loss of collagen elasticity, and subsequently to the reduction in myocardial compliance [116].
Insulin resistance significantly limits the utilization of glucose by cardiac cells, with a consequent shift towards the use of fatty acids for energy production [117]. The increased mitochondrial fatty acid uptake and β-oxidation may induce mitochondrial dysfunction and the intracellular accumulation of toxic lipids and lipid metabolites that may cause myocardial necrosis and fibrosis [118].
Hyperglycemia also induces the RAAS activation that contributes to the myocardial hypertrophy and fibrosis observed in patients affected by diabetic cardiomyopathy [119]. In patients affected by diabetes, there is an increase in the tissue expression and activity of Ang-II, which binds to angiotensin receptor-1 present on cardiomyocytes and cardiac fibroblasts and induces cell proliferation and collagen synthesis, causing cardiac hypertrophy and fibrosis. In vitro and in vivo studies showed that TNF receptor 1 signaling is necessary for the Ang-II-induced transcriptional upregulation of several fibrosis- and inflammation-related genes and the development of cardiac fibrosis [120]. The harmful effects of Ang-II are counterbalanced by the heptapeptide angiotensin-(1-7) that shows a protective role in the cardiovascular system by binding to the Mas receptor and inhibiting cardiac cells’ growth [121].
The balance between MMPs and TIMPs is essential for the regulation of ECM degradation. High glucose stimulation may lead to an imbalance in the synthesis and degradation of the ECM and collagen, thus promoting cardiac fibrosis. Experimental studies show that MMP-2 expression is downregulated in streptozotocin-induced diabetic mice, and it is associated with a reduction in collagen degradation and an increase in TGF-β and myocardial fibrosis [122]. Additionally, the blocking of the AT-1 receptor is associated with the normalization of MMP activity and the reduction in TGF-β levels and cardiac fibrosis [122].
The TGF-β signaling pathway regulates cell proliferation, differentiation and migration and gene expression, and is implicated in reparative and fibrotic processes. TGF-β transduces signaling through a type II receptor (TGFβR2) which is constitutively active on the cell surface. The binding of TGF-βs to the TGFβR2 recruits and phosphorylates type I receptor kinases (TGFβR1), which in turn phosphorylates intracellular transcriptional regulators, namely the receptor-activated Smads (R-Smads) such as Smad2 and Smad3. Activated R-Smads form complexes with the common Smad, Smad4, and translocate to the nucleus, where they regulate the transcription of target genes [123]. Smad6 and Smad7 are TGF-β antagonists, or inhibitory SMADs, that combine with active TGFβR1 and prevent Smad2/3 binding and activation. Besides the activation of the canonical Smad-dependent cascade, TGF-βs can also activate the MAPK family, including extracellular signal-regulated kinase 1/2 (ERK1/2), p38 MAPK and c-Jun amino-terminal kinase (JNK) signaling [124]. Many cell types, including cardiac fibroblasts, produce and secrete TGF-β when they are exposed to high levels of glucose [125]. TGF-β signaling is also triggered by increased levels of Ang-II, cytokines and ROS, which are able to activate local stores of TGF-β, promote the transcription and secretion of TGF-β isoforms, and induce the synthesis and externalization of TGF-β receptors on the cell surface. TGF-β signaling induces cardiac fibroblast proliferation and differentiation and ECM accumulation through the canonical Smad-dependent and Smad-independent pathways that promote the transcription of collagen I, collagen III, fibronectin, α-smooth muscle actin, TIMP-1 and growth factors such as platelet-derived growth factor, fibroblast growth factor and angiogenic growth factor [126].
Experimental studies showed that miRNAs exhibit important regulatory effects on the TGF- β signaling pathway and, conversely, that TGF-β signaling itself may influence miRNA expression and accelerate miRNA maturation [127]. In vitro studies showed that hyperglycemia induces the upregulation of miR-21 mRNA levels that, in turn, induce Smad7 downregulation and a higher phosphorylation of Smad2 and Smad3 in cardiac fibroblasts, thus promoting myocardial fibrosis [128].
Recent studies demonstrated that miR-155 is also upregulated by high glucose levels and induces the overexpression of Smad-2 [129]. Conversely, miR-15b and miR-141 exert anti-antifibrotic effects by targeting TGFβR1 and TGF-β1, respectively, thus preventing the activation of the fibrotic signaling pathway; miR-15b and miR-141 have been shown to be downregulated by hyperglycemia, with the consequent upregulation of TGFβR1 and TGF-β1.
Besides tight glycemic control, the prevention of cardiac fibrosis and remodeling is a valuable therapeutic strategy in patients with diabetic cardiomyopathy. Pharmacological drugs such as angiotensin-converting enzyme inhibitors, Ang-II receptor blockers and aldosterone antagonists have already proven their efficacy in reducing myocardial fibrosis and improving cardiac function in HF patients. Targeting the TGF- β signaling pathway and mi-RNAs may represent a promising therapeutic intervention in these patients.

6. Role of Epicardial Adipose Tissue in DCM

Epicardial adipose tissue (EAT) is a complex endocrine organ with functions that extend beyond just providing warmth and mechanical protection to the heart [130,131]. EAT is a type of visceral fat that is located between the epicardium and the visceral layer of the pericardium. It covers roughly 80% of the surface of the heart [132,133,134]. EAT presents smaller adipocytes that express uncoupling protein 1 (UCP1), typical of brown adipose tissue (BAT) [134,135]. In T2D patients, the uptake of glucose is reduced in BAT, decreasing the expression of peroxisome proliferator-activated receptor-gamma coactivator-1alpha (PGC-1α), which is a key regulator of energy metabolism [136]. Furthermore, EAT contains anti-inflammatory macrophages [137]; it has been demonstrated that EAT in T2D patients shows different gene expression, with a higher expression of genes associated with inflammation and of cytokines. This suggests that diabetes may predispose patients to detrimental cardiovascular effects by altering the inflammatory response and cytokine activity in EAT [138]. Numerous cardiovascular imaging studies have shown that patients with T2D have a significantly increased area/volume of EAT [139,140,141,142,143]. Moreover, studies using invasive EAT biopsies and histological assessments have indicated that in T2D and obesity, there is a shift in the balance of adipocytokines within the EAT. This shift favors pro-inflammatory adipocytokines over anti-inflammatory ones, leading to chronic, low-grade inflammation. This imbalance may contribute to the development of cardiovascular disease [138]. Consistently, imaging and histological studies evidence that EAT acts in a key role in the pathological development of DCM [144].

7. Conclusions

DCM represents one of the major diabetes complications. Different mechanisms have been discovered to be the causes of DCM (Figure 2). The main mechanisms have been described in this review, which led to the development of promising therapeutic targets.

Author Contributions

Conceptualization, A.G., G.B.; methodology, cell death, G.B.; fibrosis, A.G., G.B.L. and F.O. exosomes, C.B.; aquaporins, B.B. and G.P.N., epicardial adipose tissue, A.A.; writing—original draft preparation, all authors; writing—review and editing, all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Annicchiarico, A.; Barile, B.; Buccoliero, C.; Nicchia, G.P.; Brunetti, G. Alternative therapeutic strategies in diabetes management. World J. Diabetes 2024, 15, 1142–1161. [Google Scholar] [CrossRef]
  2. Faienza, M.F.; Pontrelli, P.; Brunetti, G. Type 2 diabetes and bone fragility in children and adults. World J. Diabetes 2022, 13, 900–911. [Google Scholar] [CrossRef]
  3. Tan, Y.; Zhang, Z.; Zheng, C.; Wintergerst, K.A.; Keller, B.B.; Cai, L. Mechanisms of diabetic cardiomyopathy and potential therapeutic strategies: Preclinical and clinical evidence. Nat. Rev. Cardiol. 2020, 17, 585–607. [Google Scholar] [CrossRef]
  4. Lopez, J.; Liu, Y.; Butler, J.; Del Prato, S.; Ezekowitz, J.A.; Lam, C.S.P.; Marwick, T.H.; Rosenstock, J.; Tang, W.H.W.; Perfetti, R.; et al. Racial Differences in Diabetic Cardiomyopathy: The ARISE-HF Trial. J. Am. Coll. Cardiol. 2024, 84, 233–243. [Google Scholar] [CrossRef]
  5. Brunetti, G.; D’Amato, G.; De Santis, S.; Grano, M.; Faienza, M.F. Mechanisms of altered bone remodeling in children with type 1 diabetes. World J. Diabetes 2021, 12, 997–1009. [Google Scholar] [CrossRef] [PubMed]
  6. Paolillo, S.; Marsico, F.; Prastaro, M.; Renga, F.; Esposito, L.; De Martino, F.; Di Napoli, P.; Esposito, I.; Ambrosio, A.; Ianniruberto, M.; et al. Diabetic Cardiomyopathy. Heart Fail. Clin. 2019, 15, 341–347. [Google Scholar] [CrossRef] [PubMed]
  7. Khan, S.; Ahmad, S.S.; Kamal, M.A. Diabetic cardiomyopathy: From mechanism to management in a nutshell. Endocr. Metab. Immune Disord.-Drug Targets (Former. Curr. Drug Targets-Immune Endocr. Metab. Disord.) 2021, 21, 268–281. [Google Scholar] [CrossRef]
  8. Liu, Q.; Wang, S.; Cai, L. Diabetic cardiomyopathy and its mechanisms: Role of oxidative stress and damage. J. Diabetes Investig. 2014, 5, 623–634. [Google Scholar] [CrossRef] [PubMed]
  9. Wei, J.; Zhao, Y.; Liang, H.; Du, W.; Wang, L. Preliminary evidence for the presence of multiple forms of cell death in diabetes cardiomyopathy. Acta Pharm. Sin. B 2022, 12, 1–17. [Google Scholar] [CrossRef] [PubMed]
  10. Chen, Y.; Hua, Y.; Li, X.; Arslan, I.M.; Zhang, W.; Meng, G. Distinct Types of Cell Death and the Implication in Diabetic Cardiomyopathy. Front. Pharmacol. 2020, 11, 42. [Google Scholar] [CrossRef]
  11. Dhalla, N.S.; Shah, A.K.; Tappia, P.S. Role of Oxidative Stress in Metabolic and Subcellular Abnormalities in Diabetic Cardiomyopathy. Int. J. Mol. Sci. 2020, 21, 2413. [Google Scholar] [CrossRef]
  12. Byrne, N.J.; Rajasekaran, N.S.; Abel, E.D.; Bugger, H. Therapeutic potential of targeting oxidative stress in diabetic cardiomyopathy. Free Radic. Biol. Med. 2021, 169, 317–342. [Google Scholar] [CrossRef]
  13. Giacco, F.; Brownlee, M.; Schmidt, A.M. Oxidative Stress and Diabetic Complications. Circ. Res. 2010, 107, 1058–1070. [Google Scholar] [CrossRef]
  14. Zhang, X.; Ma, X.; Zhao, M.; Zhang, B.; Chi, J.; Liu, W.; Chen, W.; Fu, Y.; Liu, Y.; Yin, X. H2 and H3 relaxin inhibit high glucose-induced apoptosis in neonatal rat ventricular myocytes. Biochimie 2015, 108, 59–67. [Google Scholar] [CrossRef] [PubMed]
  15. Khanna, S.; Singh, G.B.; Khullar, M. Nitric oxide synthases and diabetic cardiomyopathy. Nitric Oxide 2014, 43, 29–34. [Google Scholar] [CrossRef] [PubMed]
  16. Li, G.; Yang, L.; Feng, L.; Yang, J.; Li, Y.; An, J.; Li, D.; Xu, Y.; Gao, Y.; Li, J.; et al. Syringaresinol Protects against Type 1 Diabetic Cardiomyopathy by Alleviating Inflammation Responses, Cardiac Fibrosis, and Oxidative Stress. Mol. Nutr. Food Res. 2020, 64, e2000231. [Google Scholar] [CrossRef]
  17. Song, S.; Ding, Y.; Dai, G.-l.; Zhang, Y.; Xu, M.-t.; Shen, J.-r.; Chen, T.-t.; Chen, Y.; Meng, G.-l. Sirtuin 3 deficiency exacerbates diabetic cardiomyopathy via necroptosis enhancement and NLRP3 activation. Acta Pharmacol. Sin. 2020, 42, 230–241. [Google Scholar] [CrossRef] [PubMed]
  18. Dia, M.; Gomez, L.; Thibault, H.; Tessier, N.; Leon, C.; Chouabe, C.; Ducreux, S.; Gallo-Bona, N.; Tubbs, E.; Bendridi, N.; et al. Reduced reticulum-mitochondria Ca(2+) transfer is an early and reversible trigger of mitochondrial dysfunctions in diabetic cardiomyopathy. Basic. Res. Cardiol. 2020, 115, 74. [Google Scholar] [CrossRef] [PubMed]
  19. Torre, E.; Arici, M.; Lodrini, A.M.; Ferrandi, M.; Barassi, P.; Hsu, S.C.; Chang, G.J.; Boz, E.; Sala, E.; Vagni, S.; et al. SERCA2a stimulation by istaroxime improves intracellular Ca2+ handling and diastolic dysfunction in a model of diabetic cardiomyopathy. Cardiovasc. Res. 2022, 118, 1020–1032. [Google Scholar] [CrossRef]
  20. Zhao, S.M.; Wang, Y.L.; Guo, C.Y.; Chen, J.L.; Wu, Y.Q. Progressive decay of Ca2+ homeostasis in the development of diabetic cardiomyopathy. Cardiovasc. Diabetol. 2014, 13, 75. [Google Scholar] [CrossRef]
  21. Chen, Y.; Xin, Y.; Cheng, Y.; Liu, X. Mitochondria-Endoplasmic Reticulum Contacts: The Promising Regulators in Diabetic Cardiomyopathy. Oxid. Med. Cell Longev. 2022, 2022, 2531458. [Google Scholar] [CrossRef]
  22. Wang, G.; Ma, T.Y.; Huang, K.; Zhong, J.H.; Lu, S.J.; Li, J.J. Role of pyroptosis in diabetic cardiomyopathy: An updated review. Front. Endocrinol. 2023, 14, 1322907. [Google Scholar] [CrossRef] [PubMed]
  23. Bertheloot, D.; Latz, E.; Franklin, B.S. Necroptosis, pyroptosis and apoptosis: An intricate game of cell death. Cell Mol. Immunol. 2021, 18, 1106–1121. [Google Scholar] [CrossRef] [PubMed]
  24. Yu, P.; Zhang, X.; Liu, N.; Tang, L.; Peng, C.; Chen, X. Pyroptosis: Mechanisms and diseases. Signal Transduct. Target. Ther. 2021, 6, 128. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Chen, X.; Gueydan, C.; Han, J. Plasma membrane changes during programmed cell deaths. Cell Res. 2018, 28, 9–21. [Google Scholar] [CrossRef]
  26. Cai, Z.; Yuan, S.; Luan, X.; Feng, J.; Deng, L.; Zuo, Y.; Li, J. Pyroptosis-Related Inflammasome Pathway: A New Therapeutic Target for Diabetic Cardiomyopathy. Front. Pharmacol. 2022, 13, 842313. [Google Scholar] [CrossRef] [PubMed]
  27. Li, L.; Luo, W.; Qian, Y.; Zhu, W.; Qian, J.; Li, J.; Jin, Y.; Xu, X.; Liang, G. Luteolin protects against diabetic cardiomyopathy by inhibiting NF-kappaB-mediated inflammation and activating the Nrf2-mediated antioxidant responses. Phytomedicine 2019, 59, 152774. [Google Scholar] [CrossRef] [PubMed]
  28. Wei, Z.; Jing, Z.; Pinfang, K.; Chao, S.; Shaohuan, Q. Quercetin Inhibits Pyroptosis in Diabetic Cardiomyopathy through the Nrf2 Pathway. J. Diabetes Res. 2022, 2022, 9723632. [Google Scholar] [CrossRef]
  29. Zeng, C.; Wang, R.; Tan, H. Role of Pyroptosis in Cardiovascular Diseases and its Therapeutic Implications. Int. J. Biol. Sci. 2019, 15, 1345–1357. [Google Scholar] [CrossRef]
  30. Zhang, H.; Chen, X.; Zong, B.; Yuan, H.; Wang, Z.; Wei, Y.; Wang, X.; Liu, G.; Zhang, J.; Li, S.; et al. Gypenosides improve diabetic cardiomyopathy by inhibiting ROS-mediated NLRP3 inflammasome activation. J. Cell Mol. Med. 2018, 22, 4437–4448. [Google Scholar] [CrossRef]
  31. Zhang, L.; Ai, C.; Bai, M.; Niu, J.; Zhang, Z. NLRP3 Inflammasome/Pyroptosis: A Key Driving Force in Diabetic Cardiomyopathy. Int. J. Mol. Sci. 2022, 23, 10632. [Google Scholar] [CrossRef]
  32. Feng, Y.; Xu, W.; Zhang, W.; Wang, W.; Liu, T.; Zhou, X. LncRNA DCRF regulates cardiomyocyte autophagy by targeting miR-551b-5p in diabetic cardiomyopathy. Theranostics 2019, 9, 4558. [Google Scholar] [CrossRef] [PubMed]
  33. Bhat, A.A.; Goyal, A.; Thapa, R.; Kazmi, I.; Alzarea, S.I.; Singh, M.; Rohilla, S.; Saini, T.K.; Kukreti, N.; Meenakshi, D.U. Uncovering the complex role of interferon-gamma in suppressing type 2 immunity to cancer. Cytokine 2023, 171, 156376. [Google Scholar] [CrossRef] [PubMed]
  34. Dewanjee, S.; Vallamkondu, J.; Kalra, R.S.; John, A.; Reddy, P.H.; Kandimalla, R. Autophagy in the diabetic heart: A potential pharmacotherapeutic target in diabetic cardiomyopathy. Ageing Res. Rev. 2021, 68, 101338. [Google Scholar] [CrossRef] [PubMed]
  35. Bhat, A.A.; Gupta, G.; Goyal, A.; Thapa, R.; Almalki, W.H.; Kazmi, I.; Alzarea, S.I.; Kukreti, N.; Sekar, M.; Meenakshi, D.U. Unwinding circular RNA’s role in inflammatory pulmonary diseases. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2024, 397, 2567–2588. [Google Scholar] [CrossRef]
  36. Kubli, D.A.; Gustafsson, Å.B. Unbreak my heart: Targeting mitochondrial autophagy in diabetic cardiomyopathy. Antioxid. Redox Signal. 2015, 22, 1527–1544. [Google Scholar] [CrossRef]
  37. Kakoty, V.; Sarathlal, K.; Kaur, P.; Wadhwa, P.; Vishwas, S.; Khan, F.R.; Alhazmi, A.Y.M.; Almasoudi, H.H.; Gupta, G.; Chellappan, D.K. Unraveling the role of glial cell line–derived neurotrophic factor in the treatment of Parkinson’s disease. Neurol. Sci. 2024, 45, 1409–1418. [Google Scholar] [CrossRef]
  38. Madonna, R.; Moscato, S.; Cufaro, M.C.; Pieragostino, D.; Mattii, L.; Del Boccio, P.; Ghelardoni, S.; Zucchi, R.; De Caterina, R. Empagliflozin inhibits excessive autophagy through the AMPK/GSK3β signalling pathway in diabetic cardiomyopathy. Cardiovasc. Res. 2023, 119, 1175–1189. [Google Scholar] [CrossRef]
  39. Gulati, N.; Chellappan, D.K.; MacLoughlin, R.; Gupta, G.; Singh, S.K.; Oliver, B.G.; Dua, K.; Dureja, H. Advances in nano-based drug delivery systems for the management of cytokine influx-mediated inflammation in lung diseases. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2023, 397, 3695–3707. [Google Scholar] [CrossRef]
  40. Guo, Y.; Feng, X.; Wang, D.; Kang, X.; Zhang, L.; Ren, H.; Yuan, G. Long non-coding RNA: A key regulator in the pathogenesis of diabetic cardiomyopathy. Front. Cardiovasc. Med. 2021, 8, 655598. [Google Scholar] [CrossRef]
  41. Chen, S.; Liu, D.; Zhou, Z.; Qin, S. Role of long non-coding RNA H19 in the development of osteoporosis. Mol. Med. 2021, 27, 122. [Google Scholar] [CrossRef] [PubMed]
  42. Feng, L.; Yan, Q.; Xu, X. Long non-coding RNA H19 recruits NFYB to activate MBTD1 and regulate doxorubicin resistance in lymphoma cells. Mol. Biotechnol. 2023, 65, 997–1009. [Google Scholar] [CrossRef] [PubMed]
  43. Alharbi, K.S. Exploring GAS5’s impact on prostate cancer: Recent discoveries and emerging paradigms. Pathol.-Res. Pract. 2023, 251, 154851. [Google Scholar] [CrossRef]
  44. Filippova, E.A.; Fridman, M.V.; Burdennyy, A.M.; Loginov, V.I.; Pronina, I.V.; Lukina, S.S.; Dmitriev, A.A.; Braga, E.A. Long noncoding RNA GAS5 in breast cancer: Epigenetic mechanisms and biological functions. Int. J. Mol. Sci. 2021, 22, 6810. [Google Scholar] [CrossRef]
  45. Azizidoost, S.; Ghaedrahmati, F.; Anbiyaee, O.; Ahmad Ali, R.; Cheraghzadeh, M.; Farzaneh, M. Emerging roles for lncRNA-NEAT1 in colorectal cancer. Cancer Cell Int. 2022, 22, 209. [Google Scholar] [CrossRef] [PubMed]
  46. Bu, F.t.; Wang, A.; Zhu, Y.; You, H.m.; Zhang, Y.f.; Meng, X.m.; Huang, C.; Li, J. LncRNA NEAT1: Shedding light on mechanisms and opportunities in liver diseases. Liver Int. 2020, 40, 2612–2626. [Google Scholar] [CrossRef] [PubMed]
  47. You, P.; Chen, H.; Han, W.; Deng, J. miR-200a-3p overexpression alleviates diabetic cardiomyopathy injury in mice by regulating autophagy through the FOXO3/Mst1/Sirt3/AMPK axis. PeerJ 2023, 11, e15840. [Google Scholar] [CrossRef]
  48. Cui, J.; Shu, J. Circulating microRNA trafficking and regulation: Computational principles and practice. Brief. Bioinform. 2020, 21, 1313–1326. [Google Scholar] [CrossRef]
  49. Fellmann, C.; Hoffmann, T.; Sridhar, V.; Hopfgartner, B.; Muhar, M.; Roth, M.; Lai, D.Y.; Barbosa, I.A.; Kwon, J.S.; Guan, Y. An optimized microRNA backbone for effective single-copy RNAi. Cell Rep. 2013, 5, 1704–1713. [Google Scholar] [CrossRef]
  50. Break, M.; Khaled, B.; Syed, R.U.; Hussein, W.; Alqarni, S.; Magam, S.M.; Nawaz, M.; Shaikh, S.; Otaibi, A.A.; Masood, N.; et al. Noncoding RNAs as therapeutic targets in autophagy-related diabetic cardiomyopathy. Pathol.-Res. Pract. 2024, 256. [Google Scholar] [CrossRef]
  51. Costantino, S.; Paneni, F.; Lüscher, T.F.; Cosentino, F. MicroRNA profiling unveils hyperglycaemic memory in the diabetic heart. Eur. Heart J. 2016, 37, 572–576. [Google Scholar] [CrossRef]
  52. Barrera, G.; Pizzimenti, S.; Ciamporcero, E.S.; Daga, M.; Ullio, C.; Arcaro, A.; Cetrangolo, G.P.; Ferretti, C.; Dianzani, C.; Lepore, A.; et al. Role of 4-Hydroxynonenal-Protein Adducts in Human Diseases. Antioxid. Redox Signal. 2015, 22, 1681–1702. [Google Scholar] [CrossRef]
  53. Song, Z.; Wang, J.; Zhang, L. Ferroptosis: A New Mechanism in Diabetic Cardiomyopathy. Int. J. Med. Sci. 2024, 21, 612–622. [Google Scholar] [CrossRef] [PubMed]
  54. Zang, H.; Wu, W.; Qi, L.; Tan, W.; Nagarkatti, P.; Nagarkatti, M.; Wang, X.; Cui, T. Autophagy Inhibition Enables Nrf2 to Exaggerate the Progression of Diabetic Cardiomyopathy in Mice. Diabetes 2020, 69, 2720–2734. [Google Scholar] [CrossRef]
  55. Zhao, Y.; Li, Y.; Zhang, R.; Wang, F.; Wang, T.; Jiao, Y. The Role of Erastin in Ferroptosis and Its Prospects in Cancer Therapy. OncoTargets Ther. 2020, 13, 5429–5441. [Google Scholar] [CrossRef] [PubMed]
  56. Ni, T.; Huang, X.; Pan, S.; Lu, Z. Inhibition of the long non-coding RNA ZFAS1 attenuates ferroptosis by sponging miR-150-5p and activates CCND2 against diabetic cardiomyopathy. J. Cell. Mol. Med. 2021, 25, 9995–10007. [Google Scholar] [CrossRef] [PubMed]
  57. Papadopoulos, M.C.; Saadoun, S. Key roles of aquaporins in tumor biology. Biochim. Biophys. Acta (BBA)-Biomembr. 2015, 1848, 2576–2583. [Google Scholar] [CrossRef]
  58. Shangzu, Z.; Dingxiong, X.; ChengJun, M.; Yan, C.; Yangyang, L.; Zhiwei, L.; Ting, Z.; Zhiming, M.; Yiming, Z.; Liying, Z.; et al. Aquaporins: Important players in the cardiovascular pathophysiology. Pharmacol. Res. 2022, 183, 106363. [Google Scholar] [CrossRef] [PubMed]
  59. Frustaci, A.; Letizia, C.; Chimenti, C.; Verardo, R.; Alfarano, M.; Scialla, R.; Bagnato, G.; Miraldi, F.; Sansone, L.; Russo, M.A. Myocardial Aldosterone Receptor and Aquaporin 1 Up-Regulation Is Associated with Cardiomyocyte Remodeling in Human Heart Failure. J. Clin. Med. 2021, 10, 4854. [Google Scholar] [CrossRef]
  60. Li, L.; Weng, Z.; Yao, C.; Song, Y.; Ma, T. Aquaporin-1 Deficiency Protects Against Myocardial Infarction by Reducing Both Edema and Apoptosis in Mice. Sci. Rep. 2015, 5, 13807. [Google Scholar] [CrossRef]
  61. Eltobshy, S.A.G.; Messiha, R.; Metias, E.; Sarhan, M.; El-Gamal, R.; El-Shaieb, A.; Ghalwash, M. Effect of SGLT2 Inhibitor on Cardiomyopathy in a Rat Model of T2DM: Possible involvement of Cardiac Aquaporins. Tissue Cell 2023, 85, 102200. [Google Scholar] [CrossRef] [PubMed]
  62. Mani, S.; Gurusamy, N.; Ulaganathan, T.; Paluck, A.J.; Ramalingam, S.; Rajasingh, J. Therapeutic potentials of stem cell–derived exosomes in cardiovascular diseases. Exp. Biol. Med. 2023, 248, 434–444. [Google Scholar] [CrossRef] [PubMed]
  63. Mishra, A.; Bharti, P.S.; Rani, N.; Nikolajeff, F.; Kumar, S. A tale of exosomes and their implication in cancer. Biochim. Biophys. Acta (BBA)-Rev. Cancer 2023, 1878, 188908. [Google Scholar] [CrossRef]
  64. Mondal, J.; Pillarisetti, S.; Junnuthula, V.; Saha, M.; Hwang, S.R.; Park, I.-k.; Lee, Y.-k. Hybrid exosomes, exosome-like nanovesicles and engineered exosomes for therapeutic applications. J. Control. Release 2023, 353, 1127–1149. [Google Scholar] [CrossRef]
  65. Saravanan, P.B.; Kalivarathan, J.; Khan, F.; Shah, R.; Levy, M.F.; Kanak, M.A. Exosomes in transplantation: Role in allograft rejection, diagnostic biomarker, and therapeutic potential. Life Sci. 2023, 324, 121722. [Google Scholar] [CrossRef]
  66. Yao, J.; Chen, Y.; Lin, Z. Exosomes: Mediators in microenvironment of colorectal cancer. Int. J. Cancer 2023, 153, 904–917. [Google Scholar] [CrossRef] [PubMed]
  67. Ye, H.; Tan, L.; Tu, C.; Min, L. Exosomes in sarcoma: Prospects for clinical applications. Crit. Rev. Oncol./Hematol. 2023, 181, 103895. [Google Scholar] [CrossRef] [PubMed]
  68. Zhang, M.; Hu, S.; Liu, L.; Dang, P.; Liu, Y.; Sun, Z.; Qiao, B.; Wang, C. Engineered exosomes from different sources for cancer-targeted therapy. Signal Transduct. Target. Ther. 2023, 8, 124. [Google Scholar] [CrossRef] [PubMed]
  69. Xu, Z.; Chen, Y.; Ma, L.; Chen, Y.; Liu, J.; Guo, Y.; Yu, T.; Zhang, L.; Zhu, L.; Shu, Y. Role of exosomal non-coding RNAs from tumor cells and tumor-associated macrophages in the tumor microenvironment. Mol. Ther. 2022, 30, 3133–3154. [Google Scholar] [CrossRef]
  70. An, Y.; Lin, S.; Tan, X.; Zhu, S.; Nie, F.; Zhen, Y.; Gu, L.; Zhang, C.; Wang, B.; Wei, W.; et al. Exosomes from adipose-derived stem cells and application to skin wound healing. Cell Prolif. 2021, 54, e12993. [Google Scholar] [CrossRef]
  71. He, Y.; Li, Y.; Zhai, Z.; Liu, P.; Nie, L.; Zheng, Y.; Hou, J.; Huo, W.; Mao, Z.; Zhang, Z.; et al. Relationship of miRNA-126 and miRNA-122 expression with type 2 diabetes mellitus and related glucose metabolism parameters: A systematic review and meta-analysis. Exp. Ther. Med. 2022, 24, 652. [Google Scholar] [CrossRef] [PubMed]
  72. Hu, Y.; Rao, S.-S.; Wang, Z.-X.; Cao, J.; Tan, Y.-J.; Luo, J.; Li, H.-M.; Zhang, W.-S.; Chen, C.-Y.; Xie, H. Exosomes from human umbilical cord blood accelerate cutaneous wound healing through miR-21-3p-mediated promotion of angiogenesis and fibroblast function. Theranostics 2018, 8, 169–184. [Google Scholar] [CrossRef] [PubMed]
  73. Jiang, L.; Dong, H.; Cao, H.; Ji, X.; Luan, S.; Liu, J. Exosomes in Pathogenesis, Diagnosis, and Treatment of Alzheimer’s Disease. Med. Sci. Monit. 2019, 25, 3329–3335. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, J.; Jiang, F.; Jiang, Y.; Wang, Y.; Li, Z.; Shi, X.; Zhu, Y.; Wang, H.; Zhang, Z. Roles of Exosomes in Ocular Diseases. Int. J. Nanomed. 2020, 15, 10519–10538. [Google Scholar] [CrossRef]
  75. Liu, X.; Miao, J.; Wang, C.; Zhou, S.; Chen, S.; Ren, Q.; Hong, X.; Wang, Y.; Hou, F.F.; Zhou, L.; et al. Tubule-derived exosomes play a central role in fibroblast activation and kidney fibrosis. Kidney Int. 2020, 97, 1181–1195. [Google Scholar] [CrossRef]
  76. Soares Martins, T.; Trindade, D.; Vaz, M.; Campelo, I.; Almeida, M.; Trigo, G.; da Cruz e Silva, O.A.B.; Henriques, A.G. Diagnostic and therapeutic potential of exosomes in Alzheimer’s disease. J. Neurochem. 2020, 156, 162–181. [Google Scholar] [CrossRef]
  77. Zhang, X.; Wang, J.; Zhang, J.; Tan, Y.; Li, Y.; Peng, Z. Exosomes Highlight Future Directions in the Treatment of Acute Kidney Injury. Int. J. Mol. Sci. 2023, 24, 15568. [Google Scholar] [CrossRef]
  78. Liang, G.; Zhu, Y.; Ali, D.J.; Tian, T.; Xu, H.; Si, K.; Sun, B.; Chen, B.; Xiao, Z. Engineered exosomes for targeted co-delivery of miR-21 inhibitor and chemotherapeutics to reverse drug resistance in colon cancer. J. Nanobiotechnol. 2020, 18, 10. [Google Scholar] [CrossRef]
  79. Batrakova, E.V.; Kim, M.S. Using exosomes, naturally-equipped nanocarriers, for drug delivery. J. Control. Release 2015, 219, 396–405. [Google Scholar] [CrossRef] [PubMed]
  80. Liang, Y.; Duan, L.; Lu, J.; Xia, J. Engineering exosomes for targeted drug delivery. Theranostics 2021, 11, 3183–3195. [Google Scholar] [CrossRef]
  81. Rehman, F.U.; Liu, Y.; Zheng, M.; Shi, B. Exosomes based strategies for brain drug delivery. Biomaterials 2023, 293, 121949. [Google Scholar] [CrossRef] [PubMed]
  82. Fang, Y.; Ni, J.; Wang, Y.-S.; Zhao, Y.; Jiang, L.-Q.; Chen, C.; Zhang, R.-D.; Fang, X.; Wang, P.; Pan, H.-F. Exosomes as biomarkers and therapeutic delivery for autoimmune diseases: Opportunities and challenges. Autoimmun. Rev. 2023, 22, 103260. [Google Scholar] [CrossRef] [PubMed]
  83. Ren, Y.; Zhang, S.; Wang, Y.; Jacobson, D.S.; Reisdorf, R.L.; Kuroiwa, T.; Behfar, A.; Moran, S.L.; Steinmann, S.P.; Zhao, C. Effects of purified exosome product on rotator cuff tendon-bone healing in vitro and in vivo. Biomaterials 2021, 276, 121019. [Google Scholar] [CrossRef] [PubMed]
  84. Xie, R.; Zeng, X.; Yan, H.; Huang, X.; Deng, C. Effects and Mechanisms of Exosomes from Different Sources in Cerebral Ischemia. Cells 2022, 11, 3623. [Google Scholar] [CrossRef]
  85. Guo, J.; Yang, X.; Chen, J.; Wang, C.; Sun, Y.; Yan, C.; Ren, S.; Xiong, H.; Xiang, K.; Zhang, M.; et al. Exosomal miR-125b-5p derived from adipose-derived mesenchymal stem cells enhance diabetic hindlimb ischemia repair via targeting alkaline ceramidase 2. J. Nanobiotechnol. 2023, 21, 189. [Google Scholar] [CrossRef]
  86. Ilaltdinov, A.W.; Gong, Y.; Leong, D.J.; Gruson, K.I.; Zheng, D.; Fung, D.T.; Sun, L.; Sun, H.B. Advances in the development of gene therapy, noncoding RNA, and exosome-based treatments for tendinopathy. Ann. N. Y. Acad. Sci. 2020, 1490, 3–12. [Google Scholar] [CrossRef]
  87. Wang, C.; Wang, M.; Xu, T.; Zhang, X.; Lin, C.; Gao, W.; Xu, H.; Lei, B.; Mao, C. Engineering Bioactive Self-Healing Antibacterial Exosomes Hydrogel for Promoting Chronic Diabetic Wound Healing and Complete Skin Regeneration. Theranostics 2019, 9, 65–76. [Google Scholar] [CrossRef]
  88. Cui, M.; Han, Y.; Yang, J.; Li, G.; Yang, C. A narrative review of the research status of exosomes in cardiovascular disease. Ann. Palliat. Med. 2022, 11, 363–377. [Google Scholar] [CrossRef]
  89. Nandi, S.S.; Mishra, P.K. Targeting miRNA for Therapy of Juvenile and Adult Diabetic Cardiomyopathy. Adv. Exp. Med. Biol. 2018, 1056, 47–59. [Google Scholar]
  90. Sanganalmath, S.K.; Dubey, S.; Veeranki, S.; Narisetty, K.; Krishnamurthy, P. The interplay of inflammation, exosomes and Ca2+ dynamics in diabetic cardiomyopathy. Cardiovasc. Diabetol. 2023, 22, 37. [Google Scholar] [CrossRef]
  91. Tao, L.; Shi, J.; Yang, X.; Yang, L.; Hua, F. The Exosome: A New Player in Diabetic Cardiomyopathy. J. Cardiovasc. Transl. Res. 2018, 12, 62–67. [Google Scholar] [CrossRef] [PubMed]
  92. Zamani, P.; Fereydouni, N.; Butler, A.E.; Navashenaq, J.G.; Sahebkar, A. The therapeutic and diagnostic role of exosomes in cardiovascular diseases. Trends Cardiovasc. Med. 2019, 29, 313–323. [Google Scholar] [CrossRef] [PubMed]
  93. Chaturvedi, P.; Kalani, A.; Medina, I.; Familtseva, A.; Tyagi, S.C. Cardiosome mediated regulation of MMP9 in diabetic heart: Role of mir29b and mir455 in exercise. J. Cell. Mol. Med. 2015, 19, 2153–2161. [Google Scholar] [CrossRef] [PubMed]
  94. Hirai, K.; Ousaka, D.; Fukushima, Y.; Kondo, M.; Eitoku, T.; Shigemitsu, Y.; Hara, M.; Baba, K.; Iwasaki, T.; Kasahara, S.; et al. Cardiosphere-derived exosomal microRNAs for myocardial repair in pediatric dilated cardiomyopathy. Sci. Transl. Med. 2020, 12, eabb3336. [Google Scholar] [CrossRef] [PubMed]
  95. Gan, L.; Xie, D.; Liu, J.; Bond Lau, W.; Christopher, T.A.; Lopez, B.; Zhang, L.; Gao, E.; Koch, W.; Ma, X.-L.; et al. Small Extracellular Microvesicles Mediated Pathological Communications Between Dysfunctional Adipocytes and Cardiomyocytes as a Novel Mechanism Exacerbating Ischemia/Reperfusion Injury in Diabetic Mice. Circulation 2020, 141, 968–983. [Google Scholar] [CrossRef]
  96. Ni, J.; Sun, Y.; Liu, Z. The Potential of Stem Cells and Stem Cell-Derived Exosomes in Treating Cardiovascular Diseases. J. Cardiovasc. Transl. Res. 2018, 12, 51–61. [Google Scholar] [CrossRef]
  97. Yu, B.; Kim, H.W.; Gong, M.; Wang, J.; Millard, R.W.; Wang, Y.; Ashraf, M.; Xu, M. Exosomes secreted from GATA-4 overexpressing mesenchymal stem cells serve as a reservoir of anti-apoptotic microRNAs for cardioprotection. Int. J. Cardiol. 2015, 182, 349–360. [Google Scholar] [CrossRef]
  98. Huang-Doran, I.; Zhang, C.-Y.; Vidal-Puig, A. Extracellular Vesicles: Novel Mediators of Cell Communication In Metabolic Disease. Trends Endocrinol. Metab. 2017, 28, 3–18. [Google Scholar] [CrossRef]
  99. Lawson, C.; Vicencio, J.M.; Yellon, D.M.; Davidson, S.M. Microvesicles and exosomes: New players in metabolic and cardiovascular disease. J. Endocrinol. 2016, 228, R57–R71. [Google Scholar] [CrossRef]
  100. Sahoo, S.; Emanueli, C. Exosomes in Diabetic Cardiomyopathy: The Next-Generation Therapeutic Targets? Diabetes 2016, 65, 2829–2831. [Google Scholar] [CrossRef]
  101. Wang, X.; Gu, H.; Huang, W.; Peng, J.; Li, Y.; Yang, L.; Qin, D.; Essandoh, K.; Wang, Y.; Peng, T.; et al. Hsp20-Mediated Activation of Exosome Biogenesis in Cardiomyocytes Improves Cardiac Function and Angiogenesis in Diabetic Mice. Diabetes 2016, 65, 3111–3128. [Google Scholar] [CrossRef] [PubMed]
  102. Wang, X.; Huang, W.; Liu, G.; Cai, W.; Millard, R.W.; Wang, Y.; Chang, J.; Peng, T.; Fan, G.-C. Cardiomyocytes mediate anti-angiogenesis in type 2 diabetic rats through the exosomal transfer of miR-320 into endothelial cells. J. Mol. Cell. Cardiol. 2014, 74, 139–150. [Google Scholar] [CrossRef] [PubMed]
  103. Shi, R.; Zhao, L.; Cai, W.; Wei, M.; Zhou, X.; Yang, G.; Yuan, L. Maternal exosomes in diabetes contribute to the cardiac development deficiency. Biochem. Biophys. Res. Commun. 2017, 483, 602–608. [Google Scholar] [CrossRef] [PubMed]
  104. Das, S.; Halushka, M.K. Extracellular vesicle microRNA transfer in cardiovascular disease. Cardiovasc. Pathol. 2015, 24, 199–206. [Google Scholar] [CrossRef]
  105. Loyer, X.; Vion, A.-C.; Tedgui, A.; Boulanger, C.M. Microvesicles as Cell–Cell Messengers in Cardiovascular Diseases. Circ. Res. 2014, 114, 345–353. [Google Scholar] [CrossRef] [PubMed]
  106. Vrijsen, K.R.; Sluijter, J.P.G.; Schuchardt, M.W.L.; Van Balkom, B.W.M.; Noort, W.A.; Chamuleau, S.A.J.; Doevendans, P.A.F.M. Cardiomyocyte progenitor cell-derived exosomes stimulate migration of endothelial cells. J. Cell. Mol. Med. 2010, 14, 1064–1070. [Google Scholar] [CrossRef]
  107. Atlas, S.A. The Renin-Angiotensin Aldosterone System: Pathophysiological Role and Pharmacologic Inhibition. J. Manag. Care Pharm. 2007, 13 (Suppl. B), 9–20. [Google Scholar] [CrossRef]
  108. Busson, P.; Garcia, N.A.; Ontoria-Oviedo, I.; González-King, H.; Diez-Juan, A.; Sepúlveda, P. Glucose Starvation in Cardiomyocytes Enhances Exosome Secretion and Promotes Angiogenesis in Endothelial Cells. PLoS ONE 2015, 10, e0138849. [Google Scholar] [CrossRef]
  109. Huo, J.-L.; Feng, Q.; Pan, S.; Fu, W.-J.; Liu, Z.; Liu, Z. Diabetic cardiomyopathy: Early diagnostic biomarkers, pathogenetic mechanisms, and therapeutic interventions. Cell Death Discov. 2023, 9, 256. [Google Scholar] [CrossRef]
  110. Tuleta, I.; Frangogiannis, N.G. Fibrosis of the diabetic heart: Clinical significance, molecular mechanisms, and therapeutic opportunities. Adv. Drug Deliv. Rev. 2021, 176, 113904. [Google Scholar] [CrossRef]
  111. Frangogiannis, N.G. The Extracellular Matrix in Ischemic and Nonischemic Heart Failure. Circ. Res. 2019, 125, 117–146. [Google Scholar] [CrossRef] [PubMed]
  112. Frangogiannis, N.G. Pathophysiology of Myocardial Infarction. Compr. Physiol. 2015, 5, 1841–1875. [Google Scholar] [PubMed]
  113. Kurose, H. Cardiac Fibrosis and Fibroblasts. Cells 2021, 10, 1716. [Google Scholar] [CrossRef]
  114. Singh, V.P.; Baker, K.M.; Kumar, R. Activation of the intracellular renin-angiotensin system in cardiac fibroblasts by high glucose: Role in extracellular matrix production. Am. J. Physiol.-Heart Circ. Physiol. 2008, 294, H1675–H1684. [Google Scholar] [CrossRef]
  115. Lorenzo, O.; Picatoste, B.; Ares-Carrasco, S.; Ramírez, E.; Egido, J.; Tuñón, J. Potential Role of Nuclear Factor κB in Diabetic Cardiomyopathy. Mediat. Inflamm. 2011, 2011, 652097. [Google Scholar] [CrossRef] [PubMed]
  116. Aronson, D. Cross-linking of glycated collagen in the pathogenesis of arterial and myocardial stiffening of aging and diabetes. J. Hypertens. 2003, 21, 3–12. [Google Scholar] [CrossRef]
  117. Jia, G.; Whaley-Connell, A.; Sowers, J.R. Diabetic cardiomyopathy: A hyperglycaemia- and insulin-resistance-induced heart disease. Diabetologia 2017, 61, 21–28. [Google Scholar] [CrossRef]
  118. Jia, G.; DeMarco, V.G.; Sowers, J.R. Insulin resistance and hyperinsulinaemia in diabetic cardiomyopathy. Nat. Rev. Endocrinol. 2015, 12, 144–153. [Google Scholar] [CrossRef]
  119. Batista, J.P.T.; Faria, A.O.V.d.; Ribeiro, T.F.S.; Simões e Silva, A.C. The Role of Renin–Angiotensin System in Diabetic Cardiomyopathy: A Narrative Review. Life 2023, 13, 1598. [Google Scholar] [CrossRef]
  120. Duerrschmid, C.; Crawford, J.R.; Reineke, E.; Taffet, G.E.; Trial, J.; Entman, M.L.; Haudek, S.B. TNF receptor 1 signaling is critically involved in mediating angiotensin-II-induced cardiac fibrosis. J. Mol. Cell. Cardiol. 2013, 57, 59–67. [Google Scholar] [CrossRef]
  121. Iwata, M.; Cowling, R.T.; Gurantz, D.; Moore, C.; Zhang, S.; Yuan, J.X.J.; Greenberg, B.H. Angiotensin-(1–7) binds to specific receptors on cardiac fibroblasts to initiate antifibrotic and antitrophic effects. Am. J. Physiol.-Heart Circ. Physiol. 2005, 289, H2356–H2363. [Google Scholar] [CrossRef] [PubMed]
  122. Westermann, D.; Rutschow, S.; Jäger, S.; Linderer, A.; Anker, S.; Riad, A.; Unger, T.; Schultheiss, H.-P.; Pauschinger, M.; Tschöpe, C. Contributions of Inflammation and Cardiac Matrix Metalloproteinase Activity to Cardiac Failure in Diabetic Cardiomyopathy. Diabetes 2007, 56, 641–646. [Google Scholar] [CrossRef]
  123. Frangogiannis, N.G. Transforming growth factor–β in tissue fibrosis. J. Exp. Med. 2020, 217, e20190103. [Google Scholar] [CrossRef]
  124. Xu, Z.; Sun, J.; Tong, Q.; Lin, Q.; Qian, L.; Park, Y.; Zheng, Y. The Role of ERK1/2 in the Development of Diabetic Cardiomyopathy. Int. J. Mol. Sci. 2016, 17, 2001. [Google Scholar] [CrossRef]
  125. Wu, L.; Derynck, R. Essential Role of TGF-β Signaling in Glucose-Induced Cell Hypertrophy. Dev. Cell 2009, 17, 35–48. [Google Scholar] [CrossRef] [PubMed]
  126. Yue, Y.; Meng, K.; Pu, Y.; Zhang, X. Transforming growth factor beta (TGF-β) mediates cardiac fibrosis and induces diabetic cardiomyopathy. Diabetes Res. Clin. Pract. 2017, 133, 124–130. [Google Scholar] [CrossRef] [PubMed]
  127. Qin, J.; Tan, Y.; Han, Y.; Yu, L.; Liu, S.; Zhao, S.; Wan, H.; Qu, S. Interplay Between TGF-β Signaling and MicroRNA in Diabetic Cardiomyopathy. Cardiovasc. Drugs Ther. 2023; in press. [Google Scholar] [CrossRef] [PubMed]
  128. Li, Q.; Yao, Y.; Shi, S.; Zhou, M.; Zhou, Y.; Wang, M.; Chiu, J.J.; Huang, Z.; Zhang, W.; Liu, M.; et al. Inhibition of miR-21 alleviated cardiac perivascular fibrosis via repressing EndMT in T1DM. J. Cell. Mol. Med. 2019, 24, 910–920. [Google Scholar] [CrossRef]
  129. Zhang, D.; Cui, Y.; Li, B.; Luo, X.; Li, B.; Tang, Y. miR-155 regulates high glucose-induced cardiac fibrosis via the TGF-β signaling pathway. Mol. BioSyst. 2017, 13, 215–224. [Google Scholar] [CrossRef]
  130. Iacobellis, G.; Corradi, D.; Sharma, A.M. Epicardial adipose tissue: Anatomic, biomolecular and clinical relationships with the heart. Nat. Clin. Pract. Cardiovasc. Med. 2005, 2, 536–543. [Google Scholar] [CrossRef]
  131. Antonopoulos, A.S.; Antoniades, C. The role of epicardial adipose tissue in cardiac biology: Classic concepts and emerging roles. J. Physiol. 2017, 595, 3907–3917. [Google Scholar] [CrossRef] [PubMed]
  132. Aldiss, P.; Davies, G.; Woods, R.; Budge, H.; Sacks, H.S.; Symonds, M.E. ‘Browning’the cardiac and peri-vascular adipose tissues to modulate cardiovascular risk. Int. J. Cardiol. 2017, 228, 265–274. [Google Scholar] [CrossRef] [PubMed]
  133. Iacobellis, G. Epicardial adipose tissue in contemporary cardiology. Nat. Rev. Cardiol. 2022, 19, 593–606. [Google Scholar] [CrossRef] [PubMed]
  134. Lamb, H.J. Visceral and Ectopic Fat: Risk Factors for Type 2 Diabetes, Atherosclerosis, and Cardiovascular Disease; Elsevier: Amsterdam, The Netherlands, 2022. [Google Scholar]
  135. Sacks, H.S.; Fain, J.N.; Holman, B.; Cheema, P.; Chary, A.; Parks, F.; Karas, J.; Optican, R.; Bahouth, S.W.; Garrett, E. Uncoupling protein-1 and related messenger ribonucleic acids in human epicardial and other adipose tissues: Epicardial fat functioning as brown fat. J. Clin. Endocrinol. Metab. 2009, 94, 3611–3615. [Google Scholar] [CrossRef]
  136. Moreno-Santos, I.; Pérez-Belmonte, L.M.; Macías-González, M.; Mataró, M.J.; Castellano, D.; López-Garrido, M.; Porras-Martín, C.; Sánchez-Fernández, P.L.; Gómez-Doblas, J.J.; Cardona, F. Type 2 diabetes is associated with decreased PGC1α expression in epicardial adipose tissue of patients with coronary artery disease. J. Transl. Med. 2016, 14, 243. [Google Scholar] [CrossRef]
  137. Lodewijks, F.; McKinsey, T.A.; Robinson, E.L. Fat-to-heart crosstalk in health and disease. Front. Genet. 2023, 14, 990155. [Google Scholar] [CrossRef]
  138. Yang, H.; Xiong, B.; Xiong, T.; Wang, D.; Yu, W.; Liu, B.; She, Q. Identification of key genes and mechanisms of epicardial adipose tissue in patients with diabetes through bioinformatic analysis. Front. Cardiovasc. Med. 2022, 9, 927397. [Google Scholar] [CrossRef]
  139. Al-Talabany, S.; Mordi, I.; Graeme Houston, J.; Colhoun, H.M.; Weir-McCall, J.R.; Matthew, S.Z.; Looker, H.C.; Levin, D.; Belch, J.J.; Dove, F. Epicardial adipose tissue is related to arterial stiffness and inflammation in patients with cardiovascular disease and type 2 diabetes. BMC Cardiovasc. Disord. 2018, 18, 31. [Google Scholar] [CrossRef]
  140. Chowdhary, A.; Thirunavukarasu, S.; Jex, N.; Coles, L.; Bowers, C.; Sengupta, A.; Swoboda, P.; Witte, K.; Cubbon, R.; Xue, H. Coronary microvascular function and visceral adiposity in patients with normal body weight and type 2 diabetes. Obesity 2022, 30, 1079–1090. [Google Scholar] [CrossRef]
  141. Homsi, R.; Sprinkart, A.M.; Gieseke, J.; Meier-Schroers, M.; Yuecel, S.; Fischer, S.; Nadal, J.; Dabir, D.; Luetkens, J.A.; Kuetting, D.L. Cardiac magnetic resonance based evaluation of aortic stiffness and epicardial fat volume in patients with hypertension, diabetes mellitus, and myocardial infarction. Acta Radiol. 2018, 59, 65–71. [Google Scholar] [CrossRef]
  142. Huang, S.; Li, Y.; Jiang, L.; Ren, Y.; Wang, J.; Shi, K.; Yan, W.F.; Qian, W.L.; Yang, Z.G. Impact of type 2 diabetes mellitus on epicardial adipose tissue and myocardial microcirculation by MRI in postmenopausal women. J. Magn. Reson. Imaging 2022, 56, 1404–1413. [Google Scholar] [CrossRef] [PubMed]
  143. Levelt, E.; Pavlides, M.; Banerjee, R.; Mahmod, M.; Kelly, C.; Sellwood, J.; Ariga, R.; Thomas, S.; Francis, J.; Rodgers, C. Ectopic and visceral fat deposition in lean and obese patients with type 2 diabetes. J. Am. Coll. Cardiol. 2016, 68, 53–63. [Google Scholar] [CrossRef] [PubMed]
  144. Kotha, S.; Plein, S.; Greenwood, J.P.; Levelt, E. Role of epicardial adipose tissue in diabetic cardiomyopathy through the lens of cardiovascular magnetic resonance imaging–a narrative review. Ther. Adv. Endocrinol. Metab. 2024, 15, 20420188241229540. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Diabetic cardiomyopathy mechanisms. Figure was generated using Servier Medical Art, provided by Servier, licensed under Creative Commons Attribution 3.0 Unported license.
Figure 1. Diabetic cardiomyopathy mechanisms. Figure was generated using Servier Medical Art, provided by Servier, licensed under Creative Commons Attribution 3.0 Unported license.
Ijms 25 09481 g001
Figure 2. The roles of organelle damage, oxidative stress, exosomes, epicardial adipose tissue (EAT) and activated fibroblasts in DCM. Epicardial adipose tissue (EAT). The figure was generated using Servier Medical Art, provided by Servier, licensed under Creative Commons Attribution 3.0 Unported license. Angiotensin II (Ang-II), matrix metalloproteases (MMPs), tissue inhibitors of metalloproteinases (TIMPs), mitogen-activated protein kinase (MAPK), nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), microRNA (miRNA).
Figure 2. The roles of organelle damage, oxidative stress, exosomes, epicardial adipose tissue (EAT) and activated fibroblasts in DCM. Epicardial adipose tissue (EAT). The figure was generated using Servier Medical Art, provided by Servier, licensed under Creative Commons Attribution 3.0 Unported license. Angiotensin II (Ang-II), matrix metalloproteases (MMPs), tissue inhibitors of metalloproteinases (TIMPs), mitogen-activated protein kinase (MAPK), nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), microRNA (miRNA).
Ijms 25 09481 g002
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

Galeone, A.; Annicchiarico, A.; Buccoliero, C.; Barile, B.; Luciani, G.B.; Onorati, F.; Nicchia, G.P.; Brunetti, G. Diabetic Cardiomyopathy: Role of Cell Death, Exosomes, Fibrosis and Epicardial Adipose Tissue. Int. J. Mol. Sci. 2024, 25, 9481. https://doi.org/10.3390/ijms25179481

AMA Style

Galeone A, Annicchiarico A, Buccoliero C, Barile B, Luciani GB, Onorati F, Nicchia GP, Brunetti G. Diabetic Cardiomyopathy: Role of Cell Death, Exosomes, Fibrosis and Epicardial Adipose Tissue. International Journal of Molecular Sciences. 2024; 25(17):9481. https://doi.org/10.3390/ijms25179481

Chicago/Turabian Style

Galeone, Antonella, Alessia Annicchiarico, Cinzia Buccoliero, Barbara Barile, Giovanni Battista Luciani, Francesco Onorati, Grazia Paola Nicchia, and Giacomina Brunetti. 2024. "Diabetic Cardiomyopathy: Role of Cell Death, Exosomes, Fibrosis and Epicardial Adipose Tissue" International Journal of Molecular Sciences 25, no. 17: 9481. https://doi.org/10.3390/ijms25179481

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