Next Article in Journal
Genomic Exploration of Distinct Molecular Phenotypes Steering Temozolomide Resistance Development in Patient-Derived Glioblastoma Cells
Next Article in Special Issue
The Role of Circular RNA for Early Diagnosis and Improved Management of Patients with Cardiovascular Diseases
Previous Article in Journal
Whole-Exome Sequencing of 21 Families: Candidate Genes for Early-Onset High Myopia
Previous Article in Special Issue
Assessment of sST2 Behaviors to Evaluate Severity/Clinical Impact of Acute Pulmonary Embolism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unraveling the Intricate Roles of Exosomes in Cardiovascular Diseases: A Comprehensive Review of Physiological Significance and Pathological Implications

Department of Forensic Medicine, School of Basic Medicine and Biological Sciences, Soochow University, Suzhou 215123, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(21), 15677; https://doi.org/10.3390/ijms242115677
Submission received: 13 September 2023 / Revised: 21 October 2023 / Accepted: 25 October 2023 / Published: 27 October 2023

Abstract

:
Exosomes, as potent intercellular communication tools, have garnered significant attention due to their unique cargo-carrying capabilities, which enable them to influence diverse physiological and pathological functions. Extensive research has illuminated the biogenesis, secretion, and functions of exosomes. These vesicles are secreted by cells in different states, exerting either protective or harmful biological functions. Emerging evidence highlights their role in cardiovascular disease (CVD) by mediating comprehensive interactions among diverse cell types. This review delves into the significant impacts of exosomes on CVD under stress and disease conditions, including coronary artery disease (CAD), myocardial infarction, heart failure, and other cardiomyopathies. Focusing on the cellular signaling and mechanisms, we explore how exosomes mediate multifaceted interactions, particularly contributing to endothelial dysfunction, oxidative stress, and apoptosis in CVD pathogenesis. Additionally, exosomes show great promise as biomarkers, reflecting differential expressions of NcRNAs (miRNAs, lncRNAs, and circRNAs), and as therapeutic carriers for targeted CVD treatment. However, the specific regulatory mechanisms governing exosomes in CVD remain incomplete, necessitating further exploration of their characteristics and roles in various CVD-related contexts. This comprehensive review aims to provide novel insights into the biological implications of exosomes in CVD and offer innovative perspectives on the diagnosis and treatment of CVD.

1. Introduction

The cardiovascular system constitutes a dynamic network of diverse cell types that collectively facilitate blood circulation throughout the body [1,2]. In the face of the challenges posed by a complex survival environment, a complex and finely tuned system to facilitate cell-cell homeostasis mechanisms is required to ensure that the beating heart normally and efficiently pumps blood into the vessels, thus finely controlling perfusion and fluid balance. Consequently, to sustain blood supply to the heart and ensure optimal myocardial function, effective communication among various cell types within the heart, such as cardiomyocytes, smooth muscle cells, endothelial cells (ECs), fibroblasts, mast cells, immune-system related cells, and other connective tissue cells, is indispensable. This intricate communication framework plays a pivotal role in coordinating information exchange, orchestrating the operation of multi-cellular systems, and maintaining overall cardiovascular health [3,4,5,6].
Cardiovascular disease (CVD) stands as the foremost cause of global mortality, accounting for approximately 25 percent of all deaths and imposing a substantial economic burden on society and the healthcare system [7]. Manifesting as vascular and cardiac disorders, CVD is intricately linked to progressive aging and primarily involves the arteries and heart [8,9,10]. This progressive biological process significantly affects the cardiovascular system, ultimately heightening the risk of conditions such as atherosclerosis, hypertension, myocardial infarction, and stroke [10,11]. The primary clinical events underlying CVD are predominantly of vascular origin [12]. Extensive evidence underscores the important role of inflammation, particularly ECs dysfunction, in CVD pathogenesis [13]. ECs with an inflammatory phenotype cause vascular inflammation, thereby leading to cardiac dysfunction [14]. Despite advances in cardiology aimed at optimizing coronary flow restoration and improving survival rates, outcomes following ischemia/reperfusion (IR) injury remain unsatisfactory. CVD often progresses to hypertrophy and heart failure, resulting in substantial mortality rates [15,16]. Therefore, the development of early diagnostic or treatment strategies assumes paramount importance in reducing CVD-related mortality.
Exosomes, their generation, and their function mechanisms in multicellular organisms span from physiological tissue regulation to pathogenic injury and organ remodeling. The field has been stimulated by numerous studies highlighting the correlation between circulating exosomes and the risk and severity of various diseases, including CVD, diabetes, kidney disease, and cancer [17,18,19,20]. Exosomes, deriving from diverse cell types, including those within the cardiovascular system, exist in various body fluids, including blood, urine, saliva, amniotic fluid, and breast milk [21]. Recently, exosomes, which are naturally released by mammalian cells, have emerged as pivotal tools for intercellular cardiac communication [22,23,24], significantly influencing vascular integrity and CVD [25,26]. Functioning as cargo carriers, exosomes encapsulate complex substances [27], with cargo composition varying according to the originating cell type and its “health status”. Consequently, exosomes yield disparate functional outcomes upon incorporation into recipient cells. These vesicles mediate multifaceted processes and pathways, underpinning both the normal cardiovascular physiology, such as heart development and myocardial angiogenesis [28], and pathophysiological conditions, including cardiac hypertrophy [29], atherogenesis [30], heart failure [31], and hypertension [32]. Notably, cardiac exosomes can induce endothelial dysfunction [33], oxidative stress [34], and eventual apoptosis [35], forming key pathways in the pathogenesis of CVD. However, comprehensive studies elucidating exosome phenotypes in CVD patients remain scarce, and the precise role of exosomes in CVD pathophysiology remains enigmatic.
Questions surrounding exosome function have focused on understanding the fate of their components and the phenotypic and molecular alterations that they induce in recipient cells in cell culture systems. A growing body of evidence suggests that exosomes may facilitate the transfer of non-coding RNAs (NcRNAs), such as miRNAs and lncRNAs, to recipient cells, thereby triggering gene expression and signaling pathways that drive cell communication and phenotypic alterations [36,37]. Additionally, exosomes have garnered attention as potential therapeutic targets [38], drug delivery agents [39], and tools for diverse biomedical applications [40]. Their efficiency in delivery and functional versatility underscore their potential. In this review, our focus centers on exosomes and their implications in cardiac pathophysiology. Specifically, we delve into exosome size, biogenesis, secretion, and functionality. Moreover, we provide a comprehensive overview of the dual roles of exosomes in several major cardiovascular pathologies, including coronary artery disease (CAD), myocardial infarction, heart failure, and other cardiomyopathies. Lastly, we explore the diagnostic, prognostic, and therapeutic potential of exosomes in CVD.

2. Overview of Exosomes

2.1. Exosome Size and Biogenesis

Until recently, the role of extracellular vesicles (EVs), particularly exosomes, in biogenesis, and local and remote cell-cell signaling has not been fully recognized. Various types of membrane vesicles produced by cells include apoptotic bodies, microvesicles (MVs), and exosomes [41]. These vesicles can be distinguished by their size, route of origin, and contents. Exosomes, the smallest EVs released by donor cells, have diameter ranging from 30–150 nm in diameter. In contrast, MVs and apoptotic bodies are 100–1000 nm and 50–5000 nm in diameter, respectively, both belonging to the subclass of EVs [42,43,44].
Unlike apoptotic bodies, which form from vesicles of dead or dying cells, and MVs, which bud outward from the plasma membrane, the biogenesis of exosomes is a complex and tightly regulated process [45,46]. Initially, early-sorting endosomes (ESEs) are formed through plasma membrane invaginations. These ESEs, along with their cytoplasmic contents, are then internalized into late-sorting endosomes (LSEs), eventually forming multivesicular bodies (MVBs) containing numerous intraluminal vesicles (ILVs) that will become exosomes. Subsequently, MVBs can either fuse with lysosomes or autophagosomes for degradation or merge with the plasma membrane to release the enclosed ILVs as exosomes (Figure 1) [47,48,49]. Additionally, exosomes contain unique biomarkers that distinguish them from MVs and apoptotic bodies, such as tetraspanins (e.g., CD63, CD9, and CD81), TSG101 (tumor susceptibility gene 101), syntenin-1, ALIX (apoptosis-linked gene 2-interacting protein X), and ceramides. These biomarkers are enriched in exosomes and play role in the origin and biogenesis of exosomes [50,51]. However, the specific roles and functions of these proteins in exosome biogenesis require further in-depth exploration.
Initially, when exosomes were first identified in mature sheep reticulocytes in the late 1980s, they were thought to be cell fragments with no significant impact on neighboring cells [52]. However, recent evidence suggests that these nanoscale EVs are functional carriers for cargo transport [53]. Exosomes can transport a variety of molecules, including nucleic acids, proteins, lipids, amino acids, and metabolites, which reflect their cellular origin and have a profound influence on the phenotype of recipient cells [54,55]. They can shuttle between cells in an autocrine, paracrine, or endocrine manner, thereby influencing the state of the recipient cells. It is becoming increasingly clear that exosomes are important players in both the normal physiology and pathophysiology of cells.

2.2. Secretion of Exosomes

Exosome secretion is a vital cellular mechanism observed both in vivo and in vitro (Figure 2) [56,57]. This process encompasses two major modes: the endosomal sorting complex required for transport (ESCRT)-dependent and ESCRT-independent pathways [58]. ESCRT, which plays a key role in cell membrane repair, is particularly relevant to exosome biology. The release of exosomes occurs subsequent to the fusion of MVBs with the plasma membrane, with ESCRT playing an essential role in exosome formation [59,60]. Components within the ESCRT act as important regulators during the generation of MVBs and ILVs [61,62]. Notably, ESCRT depletion has been shown to inhibit exosome biogenesis [61]. The involvement of ESCRT in exosome biogenesis occurs in stages [62,63]. Initially, ESCRT-0 targets the ESEs. This is followed by activation of ESCRT-I and -II, which are involved in membrane deformation and budding, thereby isolating the parent cell cargo. ESCRT-III then drives vesicle scission. Moreover, independent of ESCRT, alternative mechanisms have also been described. In certain cellular systems, ESCRT-independent ILVs formation in MVBs requires lipid rafts and ceramide [64,65,66], and/or aggregation of tetraspanins [67,68,69]. Among tetraspanins, CD63, which is particularly enriched intracellular and is primarily confined to LSEs and lysosomes [70], plays a role in ESCRT-independent ILVs formation, thereby regulating exosome secretion.
Exosome secretion is a multifaceted process influenced by various factors, including Ras-related protein Rab GTPases, molecular motors, cytoskeletal proteins, and SNAREs [soluble Nethylmaleimide-sensitive factor (NSF) attachment protein receptors] complex proteins. Intracellular Ca2+ levels also impact exosome secretion, with elevated Ca2+ levels leading to increased exosome release. Additionally, both intracellular and extracellular pH gradients are known to influence this process [71,72]. Recent studies have shed light on the role of silencing information regulator factor 1 (SIRT1) expression in exosome release [73]. SIRT1, a histone deacetylase, plays an important role in various physiological and pathological processes, including cell metabolism, cell survival, cell senescence, DNA repair, cell proliferation, differentiation, apoptosis, and inflammation [74,75]. Reduced SIRT1 expression restricts the expression of a specific subunit of vacuolar H+ ATPase (V-ATPase), which is responsible for proper lysosomes acidification and protein degradation. This restriction can result in a decreased number and increased size of degraded MVBs prior to fusion with the plasma membrane, ultimately leading to enhanced exosome secretion from the cells [76]. Moreover, it is worth noting that mitochondrial-derived vesicles (MDVs) have been suggested as one of the sources of exosome release. In the inflammatory state, the release of MDVs is significantly increased [77,78,79,80,81,82]. During this process, mitochondrial cargo is transported to the endolysosome system for mitochondrial quality control (MQC) [83]. The Syntaxin-17-SNAP29-VAMP7 terpolymer facilitates the delivery and fusion of MDVs to endosomes/lysosomes [84]. Importantly, in addition to being transported to lysosomes for degradation, some MDVs specifically reach endosomes to form MVBs and are subsequently released as exosomes into the extracellular environment [85]. Proteomic studies have demonstrated that up to 10% of exosomes originate from mitochondria [86,87]. These findings highlight the significance of MDVs as contributors to exosome secretion and provide novel insights into the mechanisms governing exosome production and release.

2.3. Function of Exosomes

In recent years, the study of exosomes as important mediators of cell communication has emerged endlessly [88,89]. Exosomes facilitate communication with recipient cells through various mechanisms, including ligand binding directly to the receptor to activate downstream signaling, direct membrane fusion resulting in the release of exosome contents into the recipient cells, and internalization of exosomes by the recipient cells through endocytosis (including phagocytosis, macrophage, or receptor-mediated endocytosis) [90]. Exosomal proteins can promptly stimulate recipient cells upon contact, thereby regulating cellular behavior [91]. A recent unexpected discovery has unveiled that exosomes contain double-stranded DNA, mRNA, and NcRNAs (including miRNAs, IncRNAs, and circRNAs) [92,93,94,95]. Increasing evidence suggests that the proteins and NcRNAs carried by exosomes can be inserted into recipient cells, inducing transient or long-term phenotypic changes [29,91,96,97]. These contents within exosomes are encapsulated by lipid or lipoprotein complexes, providing protection from degradation during transportation. Exosomes are involved in a variety of regulatory mechanisms [98], encompassing cell signaling, cell differentiation, immune regulation, substance metabolism, gene regulation, and tumor cell growth [99,100]. These functional molecules exhibit relative stability and can reflect the cellular origin, undergoing changes in response to variations in proteins and metabolites within the cells. Exosomes derived from different cell types exhibit pleiotropic biological activities, which can be either protective or deleterious, depending largely on the origin and current state of the cells [101]. This highlights the potential of exosome contents exchange between cells as an effective mode of intercellular communication. Specific cellular components contained within exosomes possess distinct functions and targeting capabilities, suggesting a role in regulation of intercellular communication [40].
The biological significance of exosomes in diseases is still emerging, with a substantial increase in studies exploring their utility in various pathological diagnosis and treatments. Exploiting the diverse cargo of exosomes provides a multifaceted diagnostic window for disease detection and monitoring. Diseases that have garnered significant attention for the diagnostic application of exosomes include CVD [102,103], central nervous system diseases [104,105], and cancer [106,107]. Under pathological conditions, cells release a plethora of exosomes containing diverse substances that significantly contribute to the progression of diseases. It is now understood that sorting of proteins and RNA in exosomes is specifically regulated by various pathophysiological stress stimuli and disease conditions [95,108]. This allows cells to generate tailored exosomes with distinct functional characteristics that reflect their parent cell state, with stress or disease conditions manifesting in the exosome contents. Consequently, the circulation of exosomes can be used as biomarkers for diagnostic and prognostic strategies in the context of pathological processes of CVD.
Since the first isolation of exosomes from cardiomyocytes cultured in vitro in 2007, the potential role of exosomes as diagnostic biomarkers or prognostic identifiers for CVD has gained considerable attention (Figure 3). Meanwhile, exosomes have been shown to mediate communication between various cardiovascular cells, including ECs and smooth muscle cells [109], cardiomyocytes and ECs [97], and fibroblasts and cardiomyocytes [29]. Over the past decade, exosomes as carriers of miRNAs and proteins in the field of CVD have been extensively studied [110], suggesting that exosomes may play a key role in the diagnosis and prognosis of CVD. The cargo contained within exosomes and the communication among cardiovascular cells are believed to be involved in the development of CVD [111,112]. It is becoming increasingly evident that both local and long-distance intercellular communication contribute to the maintenance of normal cardiac homeostasis and responses to hypertrophic stimuli [5,113,114]. However, exosomes released by cells affected by CVD can also exert opposite effects and further exacerbate disease progression [115,116,117]. Therefore, the multifaceted functions of exosomes in cardiovascular pathophysiology hold significant promise for a wide range of applications, which will be of great significance in exploring the pathogenesis of CVD.

3. Exosomes in Cardiovascular Disease

3.1. Exosomes in Coronary Artery Disease

CAD is a prevalent disease closely linked to increased cardiovascular morbidity and mortality, imposing a substantial economic burden on the healthcare system [118,119]. Advances in our comprehension of atherosclerosis mechanisms, the development of antiplatelet and statin medications, and progress in coronary interventional procedures have significantly improved clinical outcomes for CAD patients [120]. However, individuals with CAD still face risk such as stent restenosis, adverse cardiac remodeling, and IR injury. Atherosclerosis, primarily characterized by coronary artery plaque formation, underlies CAD pathogenesis [121]. This intricate process involves oxidative stress, endothelial dysfunction, and inflammation as pivotal contributors [122,123,124,125]. Monocytes or macrophages accumulation within blood vessel walls, resulting in the production of pro-inflammatory cytokines, is a fundamental event in atherosclerosis development [126]. Notably, exosomes released by primary human monocytes can be internalized by ECs, inducing endothelial dysfunction through the TLR4 and NF-κB pathways [127]. Another study [128] reported that lipopolysaccharide (LPS)-activated exosomes released by macrophages impact gene expression and differentiation of adipocytes, potentially playing a key role in atherosclerosis.
Exosomes derived from CAD patients have been shown to inflict endothelial injury and inflammation, contributing to endothelial dysfunction [33]. Importantly, exosomes, as natural carriers of RNA, have received significant attention in recent years for their pro-inflammatory roles in early atherosclerosis stages [129,130]. NcRNAs play important roles in the development of CAD through different mechanisms (Figure 4A) [131]. Of note, miRNAs encapsulated within exosomes [132] can be secreted by a variety of cell types, including immune cells (T and B cells) [133], stem cells [134,135], peripheral blood cells (lymphocytes, monocytes, macrophages, and platelets) [136], and adipocytes. Exosomes originating from different cell types contribute significantly to inflammation [130] and participate in all stages of atherosclerosis [137]. Moreover, exosomes released by a variety of cell types and platelets carry valuable biological information about CAD onset and progression [138]. They engage in both short- and long-range intercellular communication within the cardiovascular system through the transfer of miRNAs and other mediators [139]. Recent research has revealed that cardiac exosomes can activate inflammation via the TNF-α-mediated NF-κB pathway by transporting miR-10a [138,140]. Additionally, miR-30e and miR-92a are overexpressed in the plasma exosomes of patients with coronary artery atherosclerosis and hold promise as novel diagnostic biomarkers [141,142]. There is also evidence linking the levels of miR-21-5p and miR-100-5p in exosomes to both age and the severity of CAD [143]. Clinical studies have shown elevated serum exosomal miR-208a expression in patients with acute CAD compared to healthy subjects [144]. Furthermore, exosome concentrations and the expression level of exosomal lncRNA HIF1A-AS1 in atherosclerosis patients were significantly higher than in healthy individuals suggesting potential biomarkers for prognosis prediction [145]. In these CAD patients, the mortality is significantly increased during one-year follow-up. Another study found a significant increase in plasma and cardiac exosome miRNAs 48 h after coronary artery bypass surgery [146]. These exosomes and their contained miRNAs are positively correlated with high-sensitivity cardiac troponin I (hs-cTnI), a cardiac biomarker. Moreover, compared with miRNAs and lncRNAs, circRNAs are structurally stable and have long half-lives without degradation in exosomes, making them reliable biomarkers [147,148]. Analysis of circRNA expression in plasma exosomes in patients with CAD found plasma exosomal hsa_circ_0005540, which can be used as a promising diagnostic biomarker of CAD [149].
Intriguingly, exosomes may also exert protective effects in cardiovascular health. Given their natural role as carriers of bioactive signaling molecules, exosomes have garnered attention as potential biomarkers and therapeutic agents. Studies involving THP-1 and human embryonic kidney cells have shown that HSP27-laden exosomes significantly stimulate NF-κB activation and IL-10 release, suggesting that HSP27 could be an essential exosome cargo with beneficial anti-inflammatory effects [150]. In addition, artificially benign modifications or cardioprotective medications can induce cells to produce protective exosomes. Molecularly engineered M2 macrophage-derived exosomes, further electroporated with 5-aminolevulpiate hexyl ester, have been reported to alleviate inflammation by promoting the release of anti-inflammatory cytokines [121]. Paeonol has been reported to inhibit atherosclerosis by upregulating miR-223 expression in monocyte exosomes, thereby inhibiting the STAT3 pathway [151]. NcRNAs within exosomes hold immense promise as biomarkers for CAD detection and are expected to provide crucial insight for inhibiting atherosclerosis, which may have breakthrough significance for developing new targets for the prevention and treatment of CAD.

3.2. Exosomes in Myocardial Infarction

Myocardial infarction is a detrimental cardiac event resulting from the transient or sustained occlusion of a coronary artery, leading to myocardial necrosis and loss of cardiomyocytes [152]. Acute myocardial infarction represents one of the most severe forms of coronary heart disease (CHD) [153]. The complete occlusion of blood vessels, consequent to the lack of blood flow perfusion in the myocardium, can precipitate severe heart failure, fatal arrhythmia, cardiogenic shock, and even cardiac arrest, posing a severe risk to patient lives [154]. Current clinical practice indicate that timely reperfusion therapy has significantly improved outcomes by restoring myocardial blood flow and perfusion [155,156]. However, inevitable reperfusion injury and dysregulated immune responses often lead to detrimental ventricular remodeling and increased risk of heart failure progression [157]. Therefore, early and accurate diagnosis is paramount for improving clinical outcomes and reduce mortality.
Acute myocardial infarction categorizes into ST-segment elevation myocardial infarction (STEMI) and non-ST-segment elevation myocardial infarction (NSTEMI) [153]. STEMI typically arises from coronary atherosclerosis, characterized by a prominent presence of red blood cells and red fibrin thrombi, leading to a complete disruption of coronary blood flow [158]. Notably, approximately half of the STEMI patients might develop symptoms of heart failure between 3–6 months after infarction [159]. Early diagnosis and immediate reperfusion are pivotal for mitigating the impact of myocardial infarction and infarct size, thereby reducing potential complications and heart failure after STEMI [160]. Relatively, NSTEMI is often the result of unstable plaque formations, which lead to the partial occlusion of the coronary artery, albeit with minimal blood flow [161,162]. Although Electrocardiogram (ECG) indications of ST-segment elevation serve as sensitive and specific signs of Total coronary Occlusion (TO) in STEMI patients, TO is detected in merely 25.5–34% of NSTEMI patients [153]. Patients exhibiting TO often face delayed diagnoses, restricted access to intervention, and increased complications and mortality rates. Therefore, methods for rapidly diagnosing STEMI and NSTEMI onset are clinically valuable. Such methodologies can potentially reduce the mortality rates and improve patient prognosis [161,163].
Within the myocardial infarction context, exosomes facilitate both local and remote microcommunication among recipient cells (Figure 4B) [23,164]. It is worth noting that the content [165] and quantity [146,166] of circulating exosomes undergo significant alterations during this process. This suggests that the compromised myocardium releases specific exosomes into bodily fluids. The study found that the expression of lncRNAs ENST00000556899.1 and ENST00000575985.1 in circulating exosomes in patients with acute myocardial infarction was significantly higher than that in healthy individuals, which can be used as potential biomarkers [167]. Research has also shown that serum exosomal NEAT1, miR-204, and MMP-9 serve as potent biomarkers for acute STEMI diagnosis [168]. The prevalent adoption of medical treatments and revascularization techniques have been linked to increased early survival rates in STEMI patients [169,170]. However, limited studies have ventured into forecasting biochemical indicators associated with long-term myocardial remodeling [171]. Exosomes offer a robust evaluation mechanism in this regard. Research has elucidated that exosomes sourced from STEMI-diagnosed patients who underwent coronary angioplasty notably reduce cardiomyocyte viability [172]. Comparative studies between healthy subjects and patients 3–6 months after STEMI revealed the differential regulation of 28 miRNAs (upregulated) and 49 miRNAs (downregulated). Among these, the two most underexpressed miRNAs in heart-related exosomes are hsa-miR-181a-3p and hsa-miR-874-3p [173]. Exosomal miRNAs analysis, when combined with Real-Time Three-Dimensional Spot Tracking Echocardiography (RT3D-STE), are expected to evolve into a nuanced diagnostic system for acute myocardial infarction. Such an advancement would proficiently differentiate STEMI from NSTEMI, thus ensuring detailed and precise diagnostic outcomes [153]. The circulating heart-specific exosomal miR-152-5p and miR-3681-5p are projected to be pivotal biomarkers for the early diagnosis of STEMI and NSTEMI. Therefore, exosomal miRNAs hold immense potential as biomarkers for the early diagnosis and prognositic assessment of STEMI or NSTEMI. Notwithstanding, research elucidating the relationship between STEMI or NSTEMI and exosomes remains nascent. It is imperative to discern the key mechanisms underlying cargo classification and functional regulation, which encompasses the unique packaging of miRNAs and concomitant gene expression pathways.
On the other hand, exosomes originating from normative or benign cells have been implicated in the onset and progression of cardiac outcomes as well as repair processes subsequent to myocardial infarction. Exosomes sourced from healthy individuals confer protection ischemic myocardium through the conveyance of endogenous defensive signals, including the cardioprotective protein HSP70 [174]. The cardioprotective miR-214 has been exhibited an upregulation in the heart after ischemia and is secreted by exosomes from human ECs [175,176]. Further research has found that cardiomyocyte exosomal circHIPK3 might emerge as a therapeutic target for IR [177]. The hypoxia-induced upregulation of circHIPK3 in cardiomyocyte exosomes promotes angiogenesis and limits myocardial infarction area. This phenomenon occurs partly through the miR-29a/VEGFA axis, thereby safeguarding myocardial functionality and ECs integrity after myocardial infarction, offering myocardial protection. In the peri-infarct area, the paracrine activity of exosomes derived from unharmed myocardium might orchestrate cardiomyocytes reprogramming by the transference of molecules such as RNAs and peptides, rescuing the peri-infarct area. This subsequently attenuates necrosis, inflammation, apoptosis, remodeling, and fibrosis [178]. Predominant animal studies have administered stem cell-derived exosomes via intramyocardial delivery. During myocardial infarction episodes, exosomes released by cardiac progenitor [179] or embryonic stem cells [180] modulate cardiac regeneration and remodeling. Under hypoxic conditions, cardiac progenitor cells curtail cardiac fibrosis after myocardial infarction, impede cardiomyocyte apoptosis, and increase angiogenesis or cardiac output by transporting exosomes laden with anti-fibrotic miRNAs to fibroblasts [181]. Exosomes secreted by bone marrow MSCs transport miR-29b-3p, which targets ADAMTS16, ameliorating angiogenesis and ventricular remodeling in myocardial infarction-afflicted rats [182]. Exosomal miR-25-3p sourced from MSCs alleviates myocardial infarction by targeting pro-apoptotic proteins and EZH2 [183]. In the IR injury model, exosomes derived from MSCs protect cardiac function and reduce infarct size [184]. In addition, the intravenous administration of MSC-derived exosomes reduced infarct size by an impressive 45% and suppressed systemic inflammation [185]. These findings indicate that the specific proteins, peptides, or NcRNAs encapsulated in exosomes, aimed at mitigating myocardial infarction, predominantly target the compromised region. Such intervention modifies the adverse trajectory of myocardial infarction development, underscoring the robust therapeutic efficacy of exosomes. Evaluation of the molecular heterogeneity of exosomes proves to be of paramount importance for the clinical application and for advancing the development of therapeutic vectors.

3.3. Exosomes in Heart Failure

Heart failure represents the terminal phase of various CVD. It is characterized by the inability of the heart to supply requisite blood and oxygen to peripheral tissues, fulfilling their metabolic demands [186]. This condition is marked by a series of pathological processes, including cardiomyocyte hypertrophy, cardiac fibrosis, and impaired myocardial angiogenesis [187]. CHD stands as a significant contributor to chronic heart failure [188], and cardiogenic shock is a primary symptom of acute heart failure [189]. Notwithstanding the application of contemporary therapeutic approaches, the mortality and re-admission rates for heart failure remain alarming high. Regrettably, no specific medication exists that can effectively counteract heart failure. Therefore, investigating the pathogenesis of heart failure and pinpointing molecules integral to its onset and progression are imperative for its early diagnosis, effective treatment, and the mitigation of associated complications.
Cardiac hypertrophy is a key mechanism underlying heart failure, and a series of miRNAs have been identified as regulators of cardiac hypertrophy pathogenesis [190]. However, there remains a paucity of analyses concerning the specific components and origins of miRNAs associated with cardiac hypertrophy regulation. This gap in understanding can potentially be bridged by examining exosomal miRNAs (Figure 4C). Under conditions favoring hypertrophy, exosomal miRNAs sourced from diverse cardiovascular cells target cardiomyocytes either to induce or inhibit cardiac hypertrophy. The origin of these exosomal miRNAs suggests three primary mechanisms of action: (1) Intra-cardiomyocyte communication: There is a notable reduction in exosomal miR-133a levels with cardiomyocytes located in the infarcted and surrounding areas in myocardial infarction model [191]. When released from ischemic cardiomyocytes, exosomal miR-133a may be captured by neighboring cardiomyocytes, mitigating hypertrophy by exerting inhibitory effects on cardiomyocyte necrosis and apoptosis. (2) Interplay between cardiac fibroblasts and cardiomyocytes: Cardiac fibroblasts comprise approximately 60–70% of the cardiomyocyte population. Co-culturing cardiomyocytes with fibroblast-conditioned medium has been shown to induce cardiomyocyte hypertrophy, underscoring a pivotal communicative role between the two cell types [192,193]. Cardiac fibroblast-derived exosomes are rich in miR-21-3p, which, when incorporated by cardiomyocytes, downregulates SORBS2 and PDLIM5, thereby facilitating cardiac hypertrophy [29]. (3) Communication with other cardiovascular cell types: Macrophage-derived exosomes enriched in miR-155 promote cardiac hypertrophy and fibrosis in uremic mice by activating the pro-hypertrophic FoxO3a pathway [194]. A recent investigation highlighted that PPAR-γ activation in adipocytes, originating from adipose tissue, augments the secretion of miR-200a-enriched exosomes. This, in turn, promotes cardiac hypertrophy via the mTOR pathway [195]. The intricate interplay of miRNAs, specifically their information exchange and gene regulatory roles between cardiomyocytes and other cardiovascular cells, establishes a multifaceted and dynamic network. This network crucially influences the pathogenesis and progression of cardiac hypertrophy.
Cardiac fibrosis, another important mechanism underlying heart failure, is marked by the degradation of the extracellular matrix and an accumulation of collagen, resultant from the proliferation of cardiac fibroblasts. Numerous studies have elucidated that NcRNAs play a role in the modulation of cardiac fibrosis [196,197]. Mechanically speaking, exosomal NcRNAs, hailing from a plethora of sources, contribute to the intricate regulation of various facets of cardiac fibrosis (Figure 4C). Owing to their efficacy in carrying NcRNAs, exosomes have garnered increased attention for their role in regulating cardiac fibrosis. Throughout the progression of cardiac fibrosis, an upregulation of miR-208a is observed in cardiomyocytes and their derived exosomes. This upregulation facilitates an augmented proliferation of fibroblasts and their subsequent differentiation into myofibroblasts, expediting the progression of cardiac fibrosis [198]. Exosomes that contain LINC00636 serve to inhibit MAPK1 by overexpressing miR-450a-2-3p in human pericardial fluid, thereby ameliorating cardiac fibrosis [199]. Exosome-carried miR-29a, noted for its elevated presence in the marginal ischemic zone, acts as a conduit of information between cardiomyocytes, facilitating anti-fibrotic effects and precluding ventricular dysfunction and heart failure [200]. Exosomal miR-294, derived from embryonic stem cells, has shown its capability to inhibit fibrosis after myocardial infarction, thereby preventing myocardial infarction-induced heart failure [201]. Furthermore, miR-155, found within macrophage-derived exosomes, inhibits the proliferation of cardiac fibroblasts after myocardial infarction in mice, achieving this by targeting and inhibiting SOS-1, a principal regulator of RAS activation [202]. Collectively, these findings underscore the notion that a variety of cell-derived exosomal NcRNAs can either exacerbate or mitigate cardiac fibrosis.
In addition to pathological changes in cardiac tissues, dysregulated angiogenesis, characterized by increased capillary angiogenesis and reduced energy supply to cardiomyocytes, expediently augments the trajectory towards compensatory cardiac hypertrophy and, subsequently, heart failure [203]. The proliferation of ECs stands as a cardinal contributor to this process [204]. Recent findings spotlight the role of exosomal miRNAs in modulating angiogenesis (Figure 4C). Specifically, exosomal miR-21-5-p, derived from healthy hearts, targets and represses PTEN and BCL2, concurrently activating AKT and VEGF pathways in cardiomyocytes and ECs, endorsing both cardiomyocyte proliferation and angiogenesis [205]. During myocardial infarction, M1-like macrophages release a large amount of proinflammatory exosomes that transfer miR-155 to ECs. This action culminates in the downregulation of Rac family small GTPase 1 and p21-activated kinase 2, effectively inhibiting angiogenesis [206]. Notably, the same miRNA can manifest diametrically opposite regulatory behaviors during angiogenesis in heart failure. In instance of cardiomyopathy, exosomes enriched with miR-146a, originating from ECs, significantly attenuate EC proliferation, inhibit microvascular regeneration, and deteriorate both systolic/diastolic function [207]. Contrarily, exosomal miR-146a derived from cardiosphere cells promotes myocardial angiogenesis and significantly ameliorates heart failure subsequent to acute myocardial infarction [208]. These apparent inconsistencies might pertain to the divergent phases of heart failure. While early-stage cardiac exosomes deliver cardioprotective miRNAs, those in the terminal phase may confer detrimental effects, underpinning the multifaceted nature of exosome functionality [209]. The genesis of these contrasting outcomes might be ascribable to the distinct origins of miRNAs coupled with the variable etiologies of heart failure. Further mechanistic investigations are imperative to elucidate these seeming contradictions. Thus, current evidence suggests that exosomal miRNAs exert a bidirectional influence on heart failure, modulated by a spectrum of etiologies and pathological phases.

4. Exosomes in Other Cardiomyopathy

4.1. Exosomes in Septic Cardiomyopathy

Sepsis is a severe inflammatory response syndrome caused by infection and remains a major contributor to the mortality in the intensive care unit (ICU), characterized by multi-organ dysfunction [210,211]. Cardiovascular dysfunction is a major factor that leads to sepsis-related death [212]. Numerous studies have established a link between sepsis and cardiovascular complications, such as biventricular dilatation and reduced ejection fraction [213], which can lead to CVD and increased mortality rates. Patients with septic cardiomyopathy abnormalities face a mortality rate approximately 2–3 times higher than those without myocardial dysfunction [214,215]. Previous studies have revealed that around 40% of septic patients [213] experience myocardial inhibition [216], with cytokines (IL-β, IL-6, and TNF-α) and reactive oxygen species (ROS)/reactive nitrogen species (RNS) (including nitric oxide (NO), superoxide, and peroxynitrite) contributing to myocardial depression [217,218,219]. Moreover, the presence of exosomes has been implicated in the progression of pathological inflammation (Figure 5) [220,221]. Research has shown that exosomes in sepsis elevated levels of NADPH oxidase, nitric oxide synthase (NOS), and protein disulfide isomerase (PDI) compared to healthy exosomes. In sepsis, both increased NO production and the presence of LPS can trigger the release of platelet exosomes [35]. NO-induced and septic-platelet-derived exosomes induce caspase-3 activation and apoptosis in target ECs by producing active ROS/RNS via NADPH oxidase and NOS type II. Importantly, the neutral sphingomyelinase inhibitor GW4869 is a widely used pharmacological agent to block exosome production [222,223,224]. GW4869 inhibits the ceramide-mediated inward budding of MVBs and the release of mature exosomes from MVBs [97]. Preconditioning with GW4869 can block the production of septic exosomes, significantly reducing the release of sepsis-induced exosomes and pro-inflammatory cytokines, thereby improving cardiac function and survival [225].
Furthermore, endogenous ROS production by exosomes collected from platelets of septic patients induces vascular cell apoptosis [35,226]. In the mouse model of sepsis, circulating exosomes stimulate the formation of podosome clusters in cardiac ECs, leading to increase in vitro and in vivo permeability and cardiac dysfunction [34]. Mechanically, septic exosomes contain higher levels of ROS than normal exosomes. When these ROS are delivered to ECs, they induce vascular leakage and cardiac dysfunction [34]. Interestingly, MnTBAP (MnT), a drug with superoxide dismutase (SOD) and catalase-like activity, can act as a cellular osmotic ROS scavenger, effectively reducing ROS levels within septic exosomes and diminishing their ability to promote podosome cluster formation, thereby inhibiting vascular leakage [227,228]. In addition, the MDVs released by THP-1 monocytes stimulated by LPS contain substantial amount of mitochondrial nucleic acids, proteins, and ROS. These MDV can trigger type I interferon and TNF responses in ECs, further enhancing their proinflammatory potential [229]. Interestingly, these MDVs contain a small amount of exosome-labeled proteins [229], providing compelling evidence that exosomes contribute to inflammation and the development of CVD. Therefore, these findings suggest that exosomes have the potential to induce septic cardiomyopathy, and strategies such as inhibiting exosome production or eliminating ROS within exosomes hold promise as valuable therapeutic approaches for patients with septic shock.

4.2. Exosomes in Diabetic Cardiomyopathy

Diabetes mellitus comprises a group of metabolic diseases characterized by hyperglycemia, resulting from either insufficient insulin production (type 1 diabetes) or impaired insulin action (type 2 diabetes) [230]. It is a well-established risk factor for CVD and heart failure. In the early stage of diabetes, hyperglycemia can lead to endothelial dysfunction and microvascular abnormalities [231,232]. Human diabetic cardiomyopathy is characterized by abnormal diastolic function accompanied by mild impairment of systolic function. Pathologically, diabetic cardiomyopathy is associated with cardiomyocyte hypertrophy, necrosis, apoptosis, and increased interstitial fibrosis [233]. During hyperglycemia, intermediates facilitating communication between cardiomyocytes and ECs play a pivotal role, with exosomes emerging as key mediators in this process (Figure 5) [234]. Exosomes can exert either deleterious or beneficial effects on the myocardium, depending on the physiological state of the exosome-derived cells [235,236]. In the context of diabetic cardiomyopathy, diabetic cardiomyocytes release pathogenic exosomes carrying harmful signals. These signals, such as upregulation of miR-320 and the downregulation of IGF-1, HSP20, and Ets2, are received by neighboring ECs, resulting in impaired angiogenesis and a range of cardiovascular injuries, including ventricular dysfunction, cardiac fibrosis, and cardiomyocyte apoptosis [237]. Additionally, exosomes released by diabetic cardiomyocytes have been shown to have detrimental effects on embryonic development. Studies have revealed that exosomes from the hearts of diabetic pregnant mice can induce significant developmental defects, including congenital heart defects in fetuses. Exosomes extracted from diabetic pregnant mice can traverse the maternal-fetal barrier, and exosomal miRNA sequencing analysis has revealed significant differences in the expression of multiple miRNAs, potentially contributing to a high incidence of fetal malformation [238].
Furthermore, aside from their deleterious role in pathological myocardium, exosomes also play a beneficial role in unstimulated animals, preserving the ability of cardiomyocytes to withstand the effects of diabetes and even reversing some of the damage. A recent study by Davidson et al. showed that plasma exosomes from non-diabetic mice can protect rat hearts from IR injury both in vivo and in vitro [239]. These exosomes activate various signaling pathways, including ERK1/2, P38/MAPK, and the phosphorylation of HSP27, a member of the highly cytoprotective family of HSPs, in diabetic rat cardiomyocytes. This activation promotes the interaction of HSP70 with sarcolemmal Toll-like receptor 4 (TLR4), and protects primary cardiomyocytes from IR injury in vitro [239]. While exosomes-derived from healthy rat or human have consistently exhibited a beneficial role in the heart, this role may be altered in response to stress, such as high glucose. Diabetes and hyperglycemia can impair the cardioprotective signaling role of exosomes. Moreover, HSP20, a member of the heat shock protein (HSP) family, plays an important role in intracellular defense [240]. Myocardial overexpression of HSP20 induces qualitative and quantitative changes in the cargo composition and quantity of exosomes secreted by cardiomyocytes. HSP20 directly interacts with TSG101, a protein involved in exosome production, leading to increased exosome production and functional alteration [241]. In transgenic diabetic mice with cardiac-specific overexpression of HSP20, harmful exosomes released by cardiomyocytes are converted into protective exosomes containing phosphorylated Akt, SOD1, and surviving cell, which are cell-protective proteins. These protective exosomes can be delivered to ECs, thereby restoring cardiac function in hyperglycemic conditions [242]. Importantly, blocking exosome production with GW4869 significantly offset the cardioprotective effect mediated by HSP20 in diabetic mice [241]. Hence, it is evident that exosomes from different cell states or physiological conditions may exert opposite effects on the development of diabetic cardiomyopathy. It will be interesting to continue to delve deeper into the differences in molecular changes in cardiomyocytes between diabetic patients and healthy individuals to determine what mediates the transition of exosomes to harmful or beneficial functions.

5. Conclusions and Future Directions

Exosomes present a ubiquitous presence in all biological fluids, secreted by virtually every cell type, rendering them a compelling candidate for minimally invasive liquid biopsies and offering the potential for longitudinal sampling to monitor disease progression. Exosomes possess the remarkable capacity for capturing complex molecular cargo from both extracellular and intracellular sources. Furthermore, the surface proteins on exosomes facilitate their immune capture and enrichment, rendering them valuable for comprehensive multiparameter diagnostic assays. These vesicles function as essential mediator for the transmission of a wide spectrum of signals, ranging from protective to pathological, owing to their intricate biogenesis pathways and unique loading mechanisms that adapt to distinct physiological and pathological conditions. Strategies aimed at transforming the enrichment of NcRNAs in exosome, inhibiting exosome production, or eliminating ROS from exosomes hold promise as effective means to intervene in CVD. The capacity of exosomes to deliver functional substances to diseased cells make them favorable therapeutic vectors at both fundamental and practical levels. Ongoing exploration of exosomes, either in their native form or as vehicles for drug delivery, underscores their potential as therapeutic agents. It is important to acknowledge that not all components within exosomes exhibit discernible biological functions. Therefore, a comprehensive and meticulous characterization, complemented by functional assessments, is imperative to unravel the multifaceted biological roles of exosomes in the context of CVD. The clinical translation of exosomes for CVD diagnosis and treatment holds significant promise, but it is clear that there is still a considerable distance to cover.
It is noteworthy that the study of exosomes in the field of CVD is still in its nascent stages, characterized by a paucity of relevant investigations. This scarcity may be attributed to a deficiency in knowledge related to exosomes or technical challenges. Hence, in-depth research endeavors are warranted to gain a comprehensively understanding of the biological dimensions of exosome cargo loading, targeting, delivery, as well as to discern the endogenous contents of exosomes. Currently, the mechanisms governing exosome production and the regulatory processes underpinning their roles in CVD are only partially understood. Addressing a myriad of lingering questions emerging challenges is paramount. These include the following: (1) Deciphering the regulatory mechanism governing the production of cardiac exosomes. (2) Identifying specific exosomal molecules involved in intricate intercellular communication during CVD development. (3) Establishing whether alterations in the composition and content of exosome cargo can yield insights for early diagnosis, prediction, and efficacy evaluation in CVD. (4) Evaluating the potential therapeutic utility of cardiac exosomes in CVD management. (5) Developing strategies to modulate exosome production and cargo loading post-CVD onset to mitigate disease progression. (6) Investigating approaches to modify cardiac exosomes in vivo and in vitro settings to enhance their beneficial effects while mitigating deleterious impacts. Conducting further research into cardiac exosomes will undoubtedly facilitate the discovery of clinical biomarkers for CVD, the development of novel therapeutic targets, advancements in CVD diagnosis and prognosis, and the translation of innovative into clinical practice. This progress is pivotal in alleviating the substantial burden imposed by CVD on healthcare systems and patients.

Author Contributions

Conceptualization, S.Z. (Shuai Zhang); investigation, H.X.; writing—original draft preparation, Y.W.; writing—review and editing, S.Z. (Shaohua Zhu), Y.Y., X.L. and W.L.; funding acquisition, S.Z. (Shaohua Zhu) and H.X. All authors contributed to the paper and approved the submitted draft. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 82171871) and Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. De Bruin, R.G.; Rabelink, T.J.; van Zonneveld, A.J.; van der Veer, E.P. Emerging roles for RNA-binding proteins as effectors and regulators of cardiovascular disease. Eur. Heart J. 2017, 38, 1380–1388. [Google Scholar] [CrossRef] [PubMed]
  2. Caligiuri, G. Mechanotransduction, immunoregulation, and metabolic functions of CD31 in cardiovascular pathophysiology. Cardiovasc. Res. 2019, 115, 1425–1434. [Google Scholar] [CrossRef]
  3. Ieda, M. Heart development and regeneration via cellular interaction and reprogramming. Keio J. Med. 2013, 62, 99–106. [Google Scholar] [CrossRef]
  4. Souders, C.A.; Bowers, S.L.; Baudino, T.A. Cardiac fibroblast: The renaissance cell. Circ. Res. 2009, 105, 1164–1176. [Google Scholar] [CrossRef] [PubMed]
  5. Tirziu, D.; Giordano, F.J.; Simons, M. Cell communications in the heart. Circulation 2010, 122, 928–937. [Google Scholar] [CrossRef]
  6. Zhao, W.; Zheng, X.L.; Zhao, S.P. Exosome and its roles in cardiovascular diseases. Heart Fail. Rev. 2015, 20, 337–348. [Google Scholar] [CrossRef]
  7. Go, A.S.; Mozaffarian, D.; Roger, V.L.; Benjamin, E.J.; Berry, J.D.; Blaha, M.J.; Dai, S.; Ford, E.S.; Fox, C.S.; Franco, S.; et al. Executive summary: Heart disease and stroke statistics—2014 update: A report from the American Heart Association. Circulation 2014, 129, 399–410. [Google Scholar] [CrossRef]
  8. Cohn, J.N. Identifying the risk and preventing the consequences of cardiovascular disease. Heart Lung Circ. 2013, 22, 512–516. [Google Scholar] [CrossRef]
  9. Jokinen, E. Obesity and cardiovascular disease. Minerva Pediatr. 2015, 67, 25–32. [Google Scholar] [PubMed]
  10. North, B.J.; Sinclair, D.A. The intersection between aging and cardiovascular disease. Circ. Res. 2012, 110, 1097–1108. [Google Scholar] [CrossRef]
  11. Lakatta, E.G.; Levy, D. Arterial and cardiac aging: Major shareholders in cardiovascular disease enterprises: Part I: Aging arteries: A “set up” for vascular disease. Circulation 2003, 107, 139–146. [Google Scholar] [CrossRef] [PubMed]
  12. Cohn, J.N. New approaches to screening for vascular and cardiac risk. Am. J. Hypertens. 2001, 14 Pt 2, 218s–220s. [Google Scholar] [CrossRef]
  13. Carter, A.M. Complement activation: An emerging player in the pathogenesis of cardiovascular disease. Scientifica 2012, 2012, 402783. [Google Scholar] [CrossRef] [PubMed]
  14. Ooi, B.K.; Chan, K.G.; Goh, B.H.; Yap, W.H. The Role of Natural Products in Targeting Cardiovascular Diseases via Nrf2 Pathway: Novel Molecular Mechanisms and Therapeutic Approaches. Front. Pharmacol. 2018, 9, 1308. [Google Scholar] [CrossRef]
  15. Zhou, M.; Yu, Y.; Luo, X.; Wang, J.; Lan, X.; Liu, P.; Feng, Y.; Jian, W. Myocardial Ischemia-Reperfusion Injury: Therapeutics from a Mitochondria-Centric Perspective. Cardiology 2021, 146, 781–792. [Google Scholar] [CrossRef] [PubMed]
  16. Mokhtari-Zaer, A.; Marefati, N.; Atkin, S.L.; Butler, A.E.; Sahebkar, A. The protective role of curcumin in myocardial ischemia-reperfusion injury. J. Cell. Physiol. 2018, 234, 214–222. [Google Scholar] [CrossRef]
  17. Jadli, A.S.; Parasor, A.; Gomes, K.P.; Shandilya, R.; Patel, V.B. Exosomes in Cardiovascular Diseases: Pathological Potential of Nano-Messenger. Front. Cardiovasc. Med. 2021, 8, 767488. [Google Scholar] [CrossRef]
  18. Castaño, C.; Novials, A.; Párrizas, M. Exosomes and diabetes. Diabetes/Metab. Res. Rev. 2019, 35, e3107. [Google Scholar] [CrossRef]
  19. Ding, H.; Li, L.X.; Harris, P.C.; Yang, J.; Li, X. Extracellular vesicles and exosomes generated from cystic renal epithelial cells promote cyst growth in autosomal dominant polycystic kidney disease. Nat. Commun. 2021, 12, 4548. [Google Scholar] [CrossRef]
  20. Paskeh, M.D.A.; Entezari, M.; Mirzaei, S.; Zabolian, A.; Saleki, H.; Naghdi, M.J.; Sabet, S.; Khoshbakht, M.A.; Hashemi, M.; Hushmandi, K.; et al. Emerging role of exosomes in cancer progression and tumor microenvironment remodeling. J. Hematol. Oncol. 2022, 15, 83. [Google Scholar] [CrossRef]
  21. Yavuz, B.; Darici, H.; Zorba Yildiz, A.P.; Abamor, E.; Topuzoğullari, M.; Bağirova, M.; Allahverdiyev, A.; Karaoz, E. Formulating and Characterizing an Exosome-based Dopamine Carrier System. J. Vis. Exp. 2022, 182, e63624. [Google Scholar] [CrossRef]
  22. Liao, Z.; Chen, Y.; Duan, C.; Zhu, K.; Huang, R.; Zhao, H.; Hintze, M.; Pu, Q.; Yuan, Z.; Lv, L.; et al. Cardiac telocytes inhibit cardiac microvascular endothelial cell apoptosis through exosomal miRNA-21-5p-targeted cdip1 silencing to improve angiogenesis following myocardial infarction. Theranostics 2021, 11, 268–291. [Google Scholar] [CrossRef]
  23. Henning, R.J. Cardiovascular Exosomes and MicroRNAs in Cardiovascular Physiology and Pathophysiology. J. Cardiovasc. Transl. Res. 2021, 14, 195–212. [Google Scholar] [CrossRef]
  24. Zhao, Z.; Guo, N.; Chen, W.; Wang, Z. Leveraging Extracellular Non-coding RNAs to Diagnose and Treat Heart Diseases. J. Cardiovasc. Transl. Res. 2022, 15, 456–468. [Google Scholar] [CrossRef] [PubMed]
  25. Kishore, R.; Garikipati, V.N.S.; Gumpert, A. Tiny Shuttles for Information Transfer: Exosomes in Cardiac Health and Disease. J. Cardiovasc. Transl. Res. 2016, 9, 169–175. [Google Scholar] [CrossRef] [PubMed]
  26. 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]
  27. Reiss, A.B.; Ahmed, S.; Johnson, M.; Saeedullah, U.; De Leon, J. Exosomes in Cardiovascular Disease: From Mechanism to Therapeutic Target. Metabolites 2023, 13, 479. [Google Scholar] [CrossRef]
  28. Ribeiro, M.F.; Zhu, H.; Millard, R.W.; Fan, G.C. Exosomes Function in Pro- and Anti-Angiogenesis. Curr. Angiogenesis 2013, 2, 54–59. [Google Scholar] [CrossRef]
  29. Bang, C.; Batkai, S.; Dangwal, S.; Gupta, S.K.; Foinquinos, A.; Holzmann, A.; Just, A.; Remke, J.; Zimmer, K.; Zeug, A.; et al. Cardiac fibroblast-derived microRNA passenger strand-enriched exosomes mediate cardiomyocyte hypertrophy. J. Clin. Investig. 2014, 124, 2136–2146. [Google Scholar] [CrossRef]
  30. Gioia, M.; Vindigni, G.; Testa, B.; Raniolo, S.; Fasciglione, G.F.; Coletta, M.; Biocca, S. Membrane Cholesterol Modulates LOX-1 Shedding in Endothelial Cells. PLoS ONE 2015, 10, e0141270. [Google Scholar] [CrossRef] [PubMed]
  31. Campbell, C.R.; Berman, A.E.; Weintraub, N.L.; Tang, Y.L. Electrical stimulation to optimize cardioprotective exosomes from cardiac stem cells. Med. Hypotheses 2016, 88, 6–9. [Google Scholar] [CrossRef] [PubMed]
  32. Van Balkom, B.W.; Eisele, A.S.; Pegtel, D.M.; Bervoets, S.; Verhaar, M.C. Quantitative and qualitative analysis of small RNAs in human endothelial cells and exosomes provides insights into localized RNA processing, degradation and sorting. J. Extracell. Vesicles 2015, 4, 26760. [Google Scholar] [CrossRef]
  33. Zhang, P.; Liang, T.; Wang, X.; Wu, T.; Xie, Z.; Yu, Y.; Yu, H. Serum-Derived Exosomes from Patients with Coronary Artery Disease Induce Endothelial Injury and Inflammation in Human Umbilical Vein Endothelial Cells. Int. Heart J. 2021, 62, 396–406. [Google Scholar] [CrossRef]
  34. Mu, X.; Wang, X.; Huang, W.; Wang, R.T.; Essandoh, K.; Li, Y.; Pugh, A.M.; Peng, J.; Deng, S.; Wang, Y.; et al. Circulating Exosomes Isolated from Septic Mice Induce Cardiovascular Hyperpermeability through Promoting Podosome Cluster Formation. Shock 2018, 49, 429–441. [Google Scholar] [CrossRef] [PubMed]
  35. Gambim, M.H.; do Carmo Ade, O.; Marti, L.; Veríssimo-Filho, S.; Lopes, L.R.; Janiszewski, M. Platelet-derived exosomes induce endothelial cell apoptosis through peroxynitrite generation: Experimental evidence for a novel mechanism of septic vascular dysfunction. Crit. Care 2007, 11, R107. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, W.; Yue, C.; Gao, S.; Li, S.; Zhou, J.; Chen, J.; Fu, J.; Sun, W.; Hua, C. Promising Roles of Exosomal microRNAs in Systemic Lupus Erythematosus. Front. Immunol. 2021, 12, 757096. [Google Scholar] [CrossRef] [PubMed]
  37. Ormazabal, V.; Nair, S.; Carrión, F.; McIntyre, H.D.; Salomon, C. The link between gestational diabetes and cardiovascular diseases: Potential role of extracellular vesicles. Cardiovasc. Diabetol. 2022, 21, 174. [Google Scholar] [CrossRef]
  38. Rasaei, R.; Tyagi, A.; Rasaei, S.; Lee, S.J.; Yang, S.R.; Kim, K.S.; Ramakrishna, S.; Hong, S.H. Human pluripotent stem cell-derived macrophages and macrophage-derived exosomes: Therapeutic potential in pulmonary fibrosis. Stem Cell Res. Ther. 2022, 13, 433. [Google Scholar] [CrossRef]
  39. Li, Y.J.; Wu, J.Y.; Wang, J.M.; Hu, X.B.; Cai, J.X.; Xiang, D.X. Gemcitabine loaded autologous exosomes for effective and safe chemotherapy of pancreatic cancer. Acta Biomater. 2020, 101, 519–530. [Google Scholar] [CrossRef]
  40. Kalluri, R.; LeBleu, V.S. The biology, function, and biomedical applications of exosomes. Science 2020, 367, 6478. [Google Scholar] [CrossRef]
  41. Sluijter, J.P.; Verhage, V.; Deddens, J.C.; van den Akker, F.; Doevendans, P.A. Microvesicles and exosomes for intracardiac communication. Cardiovasc. Res. 2014, 102, 302–311. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, S.; Guo, M.; Guo, T.; Yang, M.; Cheng, J.; Cui, C.; Kang, J.; Wang, J.; Nian, Y.; Ma, W.; et al. DAL-1/4.1B promotes the uptake of exosomes in lung cancer cells via Heparan Sulfate Proteoglycan 2 (HSPG2). Mol. Cell. Biochem. 2022, 477, 241–254. [Google Scholar] [CrossRef] [PubMed]
  43. Muralidharan-Chari, V.; Clancy, J.; Plou, C.; Romao, M.; Chavrier, P.; Raposo, G.; D’Souza-Schorey, C. ARF6-regulated shedding of tumor cell-derived plasma membrane microvesicles. Curr. Biol. CB 2009, 19, 1875–1885. [Google Scholar] [CrossRef] [PubMed]
  44. Mathivanan, S.; Ji, H.; Simpson, R.J. Exosomes: Extracellular organelles important in intercellular communication. J. Proteom. 2010, 73, 1907–1920. [Google Scholar] [CrossRef] [PubMed]
  45. Raposo, G.; Stoorvogel, W. Extracellular vesicles: Exosomes, microvesicles, and friends. J. Cell Biol. 2013, 200, 373–383. [Google Scholar] [CrossRef] [PubMed]
  46. Ibrahim, A.; Marbán, E. Exosomes: Fundamental Biology and Roles in Cardiovascular Physiology. Annu. Rev. Physiol. 2016, 78, 67–83. [Google Scholar] [CrossRef]
  47. Zomer, A.; Vendrig, T.; Hopmans, E.S.; van Eijndhoven, M.; Middeldorp, J.M.; Pegtel, D.M. Exosomes: Fit to deliver small RNA. Commun. Integr. Biol. 2010, 3, 447–450. [Google Scholar] [CrossRef] [PubMed]
  48. Van Niel, G.; D’Angelo, G.; Raposo, G. Shedding light on the cell biology of extracellular vesicles. Nat. Rev. Mol. Cell Biol. 2018, 19, 213–228. [Google Scholar] [CrossRef]
  49. Kahlert, C.; Kalluri, R. Exosomes in tumor microenvironment influence cancer progression and metastasis. J. Mol. Med. 2013, 91, 431–437. [Google Scholar] [CrossRef]
  50. Macías, M.; Alegre, E.; Díaz-Lagares, A.; Patiño, A.; Pérez-Gracia, J.L.; Sanmamed, M.; López-López, R.; Varo, N.; González, A. Liquid Biopsy: From Basic Research to Clinical Practice. Adv. Clin. Chem. 2018, 83, 73–119. [Google Scholar] [CrossRef]
  51. De Abreu, R.C.; Fernandes, H.; da Costa Martins, P.A.; Sahoo, S.; Emanueli, C.; Ferreira, L. Native and bioengineered extracellular vesicles for cardiovascular therapeutics. Nat. Rev. Cardiol. 2020, 17, 685–697. [Google Scholar] [CrossRef] [PubMed]
  52. Johnstone, R.M.; Adam, M.; Hammond, J.R.; Orr, L.; Turbide, C. Vesicle formation during reticulocyte maturation. Association of plasma membrane activities with released vesicles (exosomes). J. Biol. Chem. 1987, 262, 9412–9420. [Google Scholar] [CrossRef] [PubMed]
  53. Hajinejad, M.; Sahab-Negah, S. Neuroinflammation: The next target of exosomal microRNAs derived from mesenchymal stem cells in the context of neurological disorders. J. Cell. Physiol. 2021, 236, 8070–8081. [Google Scholar] [CrossRef]
  54. Yi, X.; Chen, J.; Huang, D.; Feng, S.; Yang, T.; Li, Z.; Wang, X.; Zhao, M.; Wu, J.; Zhong, T. Current perspectives on clinical use of exosomes as novel biomarkers for cancer diagnosis. Front. Oncol. 2022, 12, 966981. [Google Scholar] [CrossRef] [PubMed]
  55. Huang, J.; Cao, H.; Cui, B.; Ma, X.; Gao, L.; Yu, C.; Shen, F.; Yang, X.; Liu, N.; Qiu, A.; et al. Mesenchymal Stem Cells-Derived Exosomes Ameliorate Ischemia/Reperfusion Induced Acute Kidney Injury in a Porcine Model. Front. Cell Dev. Biol. 2022, 10, 899869. [Google Scholar] [CrossRef] [PubMed]
  56. Singh, R.; Pochampally, R.; Watabe, K.; Lu, Z.; Mo, Y.Y. Exosome-mediated transfer of miR-10b promotes cell invasion in breast cancer. Mol. Cancer 2014, 13, 256. [Google Scholar] [CrossRef]
  57. Arita, T.; Ichikawa, D.; Konishi, H.; Komatsu, S.; Shiozaki, A.; Ogino, S.; Fujita, Y.; Hiramoto, H.; Hamada, J.; Shoda, K.; et al. Tumor exosome-mediated promotion of adhesion to mesothelial cells in gastric cancer cells. Oncotarget 2016, 7, 56855–56863. [Google Scholar] [CrossRef]
  58. Babst, M. MVB vesicle formation: ESCRT-dependent, ESCRT-independent and everything in between. Curr. Opin. Cell Biol. 2011, 23, 452–457. [Google Scholar] [CrossRef] [PubMed]
  59. Hurley, J.H. ESCRT complexes and the biogenesis of multivesicular bodies. Curr. Opin. Cell Biol. 2008, 20, 4–11. [Google Scholar] [CrossRef] [PubMed]
  60. Williams, R.L.; Urbé, S. The emerging shape of the ESCRT machinery. Nat. Rev. Mol. Cell Biol. 2007, 8, 355–368. [Google Scholar] [CrossRef] [PubMed]
  61. Colombo, M.; Moita, C.; van Niel, G.; Kowal, J.; Vigneron, J.; Benaroch, P.; Manel, N.; Moita, L.F.; Théry, C.; Raposo, G. Analysis of ESCRT functions in exosome biogenesis, composition and secretion highlights the heterogeneity of extracellular vesicles. J. Cell Sci. 2013, 126 Pt 24, 5553–5565. [Google Scholar] [CrossRef] [PubMed]
  62. Simons, M.; Raposo, G. Exosomes--vesicular carriers for intercellular communication. Curr. Opin. Cell Biol. 2009, 21, 575–581. [Google Scholar] [CrossRef] [PubMed]
  63. Bang, C.; Thum, T. Exosomes: New players in cell-cell communication. Int. J. Biochem. Cell Biol. 2012, 44, 2060–2064. [Google Scholar] [CrossRef] [PubMed]
  64. Van Niel, G.; Charrin, S.; Simoes, S.; Romao, M.; Rochin, L.; Saftig, P.; Marks, M.S.; Rubinstein, E.; Raposo, G. The tetraspanin CD63 regulates ESCRT-independent and -dependent endosomal sorting during melanogenesis. Dev. Cell 2011, 21, 708–721. [Google Scholar] [CrossRef] [PubMed]
  65. De Gassart, A.; Geminard, C.; Fevrier, B.; Raposo, G.; Vidal, M. Lipid raft-associated protein sorting in exosomes. Blood 2003, 102, 4336–4344. [Google Scholar] [CrossRef]
  66. Trajkovic, K.; Hsu, C.; Chiantia, S.; Rajendran, L.; Wenzel, D.; Wieland, F.; Schwille, P.; Brügger, B.; Simons, M. Ceramide triggers budding of exosome vesicles into multivesicular endosomes. Science 2008, 319, 1244–1247. [Google Scholar] [CrossRef]
  67. Fang, Y.; Wu, N.; Gan, X.; Yan, W.; Morrell, J.C.; Gould, S.J. Higher-order oligomerization targets plasma membrane proteins and HIV gag to exosomes. PLoS Biol. 2007, 5, e158. [Google Scholar] [CrossRef]
  68. Vidal, M.; Mangeat, P.; Hoekstra, D. Aggregation reroutes molecules from a recycling to a vesicle-mediated secretion pathway during reticulocyte maturation. J. Cell Sci. 1997, 110 Pt 16, 1867–1877. [Google Scholar] [CrossRef]
  69. Charrin, S.; le Naour, F.; Silvie, O.; Milhiet, P.E.; Boucheix, C.; Rubinstein, E. Lateral organization of membrane proteins: Tetraspanins spin their web. Biochem. J. 2009, 420, 133–154. [Google Scholar] [CrossRef]
  70. Pols, M.S.; Klumperman, J. Trafficking and function of the tetraspanin CD63. Exp. Cell Res. 2009, 315, 1584–1592. [Google Scholar] [CrossRef]
  71. Parolini, I.; Federici, C.; Raggi, C.; Lugini, L.; Palleschi, S.; De Milito, A.; Coscia, C.; Iessi, E.; Logozzi, M.; Molinari, A.; et al. Microenvironmental pH is a key factor for exosome traffic in tumor cells. J. Biol. Chem. 2009, 284, 34211–34222. [Google Scholar] [CrossRef] [PubMed]
  72. Savina, A.; Furlán, M.; Vidal, M.; Colombo, M.I. Exosome release is regulated by a calcium-dependent mechanism in K562 cells. J. Biol. Chem. 2003, 278, 20083–20090. [Google Scholar] [CrossRef] [PubMed]
  73. McAndrews, K.M.; LeBleu, V.S.; Kalluri, R. SIRT1 Regulates Lysosome Function and Exosome Secretion. Dev. Cell 2019, 49, 302–303. [Google Scholar] [CrossRef] [PubMed]
  74. Kitada, M.; Ogura, Y.; Koya, D. The protective role of Sirt1 in vascular tissue: Its relationship to vascular aging and atherosclerosis. Aging 2016, 8, 2290–2307. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, G.; Wang, F.; Ren, J.; Qiu, Y.; Zhang, W.; Gao, S.; Yang, D.; Wang, Z.; Liang, A.; Gao, Z.; et al. SIRT1 Involved in the Regulation of Alternative Splicing Affects the DNA Damage Response in Neural Stem Cells. Cell. Physiol. Biochem. Int. J. Exp. Cell. Physiol. Biochem. Pharmacol. 2018, 48, 657–669. [Google Scholar] [CrossRef]
  76. Latifkar, A.; Ling, L.; Hingorani, A.; Johansen, E.; Clement, A.; Zhang, X.; Hartman, J.; Fischbach, C.; Lin, H.; Cerione, R.A.; et al. Loss of Sirtuin 1 Alters the Secretome of Breast Cancer Cells by Impairing Lysosomal Integrity. Dev. Cell 2019, 49, 393–408.e397. [Google Scholar] [CrossRef]
  77. Krysko, D.V.; Agostinis, P.; Krysko, O.; Garg, A.D.; Bachert, C.; Lambrecht, B.N.; Vandenabeele, P. Emerging role of damage-associated molecular patterns derived from mitochondria in inflammation. Trends Immunol. 2011, 32, 157–164. [Google Scholar] [CrossRef]
  78. Soubannier, V.; McLelland, G.L.; Zunino, R.; Braschi, E.; Rippstein, P.; Fon, E.A.; McBride, H.M. A vesicular transport pathway shuttles cargo from mitochondria to lysosomes. Curr. Biol. 2012, 22, 135–141. [Google Scholar] [CrossRef]
  79. McLelland, G.L.; Soubannier, V.; Chen, C.X.; McBride, H.M.; Fon, E.A. Parkin and PINK1 function in a vesicular trafficking pathway regulating mitochondrial quality control. EMBO J. 2014, 33, 282–295. [Google Scholar] [CrossRef]
  80. Bernimoulin, M.; Waters, E.K.; Foy, M.; Steele, B.M.; Sullivan, M.; Falet, H.; Walsh, M.T.; Barteneva, N.; Geng, J.G.; Hartwig, J.H.; et al. Differential stimulation of monocytic cells results in distinct populations of microparticles. J. Thromb. Haemost. 2009, 7, 1019–1028. [Google Scholar] [CrossRef]
  81. Boudreau, L.H.; Duchez, A.C.; Cloutier, N.; Soulet, D.; Martin, N.; Bollinger, J.; Paré, A.; Rousseau, M.; Naika, G.S.; Lévesque, T.; et al. Platelets release mitochondria serving as substrate for bactericidal group IIA-secreted phospholipase A2 to promote inflammation. Blood 2014, 124, 2173–2183. [Google Scholar] [CrossRef] [PubMed]
  82. Garcia-Martinez, I.; Santoro, N.; Chen, Y.; Hoque, R.; Ouyang, X.; Caprio, S.; Shlomchik, M.J.; Coffman, R.L.; Candia, A.; Mehal, W.Z. Hepatocyte mitochondrial DNA drives nonalcoholic steatohepatitis by activation of TLR9. J. Clin. Investig. 2016, 126, 859–864. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, D.; Dong, Z.; Wang, J.; Tao, Y.; Sun, X.; Yao, X. The existence and function of mitochondrial component in extracellular vesicles. Mitochondrion 2020, 54, 122–127. [Google Scholar] [CrossRef] [PubMed]
  84. McLelland, G.L.; Lee, S.A.; McBride, H.M.; Fon, E.A. Syntaxin-17 delivers PINK1/parkin-dependent mitochondrial vesicles to the endolysosomal system. J. Cell Biol. 2016, 214, 275–291. [Google Scholar] [CrossRef] [PubMed]
  85. Picca, A.; Guerra, F.; Calvani, R.; Bucci, C.; Lo Monaco, M.R.; Bentivoglio, A.R.; Coelho-Júnior, H.J.; Landi, F.; Bernabei, R.; Marzetti, E. Mitochondrial Dysfunction and Aging: Insights from the Analysis of Extracellular Vesicles. Int. J. Mol. Sci. 2019, 20, 805. [Google Scholar] [CrossRef]
  86. Choi, D.S.; Kim, D.K.; Kim, Y.K.; Gho, Y.S. Proteomics, transcriptomics and lipidomics of exosomes and ectosomes. Proteomics 2013, 13, 1554–1571. [Google Scholar] [CrossRef]
  87. Vasam, G.; Nadeau, R.; Cadete, V.J.J.; Lavallée-Adam, M.; Menzies, K.J.; Burelle, Y. Proteomics characterization of mitochondrial-derived vesicles under oxidative stress. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2021, 35, e21278. [Google Scholar] [CrossRef]
  88. Jalaludin, I.; Lubman, D.M.; Kim, J. A guide to mass spectrometric analysis of extracellular vesicle proteins for biomarker discovery. Mass Spectrom. Rev. 2023, 42, 844–872. [Google Scholar] [CrossRef] [PubMed]
  89. Sun, Y.; Sun, F.; Jin, J.; Xu, W.; Qian, H. Exosomal LncRNAs in Gastrointestinal Cancer: Biological Functions and Emerging Clinical Applications. Cancers 2023, 15, 959. [Google Scholar] [CrossRef]
  90. Turturici, G.; Tinnirello, R.; Sconzo, G.; Geraci, F. Extracellular membrane vesicles as a mechanism of cell-to-cell communication: Advantages and disadvantages. Am. J. Physiol. Cell Physiol. 2014, 306, C621–C633. [Google Scholar] [CrossRef]
  91. Record, M.; Subra, C.; Silvente-Poirot, S.; Poirot, M. Exosomes as intercellular signalosomes and pharmacological effectors. Biochem. Pharmacol. 2011, 81, 1171–1182. [Google Scholar] [CrossRef]
  92. Thakur, B.K.; Zhang, H.; Becker, A.; Matei, I.; Huang, Y.; Costa-Silva, B.; Zheng, Y.; Hoshino, A.; Brazier, H.; Xiang, J.; et al. Double-stranded DNA in exosomes: A novel biomarker in cancer detection. Cell Res. 2014, 24, 766–769. [Google Scholar] [CrossRef]
  93. Gezer, U.; Özgür, E.; Cetinkaya, M.; Isin, M.; Dalay, N. Long non-coding RNAs with low expression levels in cells are enriched in secreted exosomes. Cell Biol. Int. 2014, 38, 1076–1079. [Google Scholar] [CrossRef] [PubMed]
  94. Essandoh, K.; Fan, G.C. Role of extracellular and intracellular microRNAs in sepsis. Biochim. Biophys. Acta 2014, 1842, 2155–2162. [Google Scholar] [CrossRef]
  95. Villarroya-Beltri, C.; Baixauli, F.; Gutiérrez-Vázquez, C.; Sánchez-Madrid, F.; Mittelbrunn, M. Sorting it out: Regulation of exosome loading. Semin. Cancer Biol. 2014, 28, 3–13. [Google Scholar] [CrossRef]
  96. Valadi, H.; Ekström, K.; Bossios, A.; Sjöstrand, M.; Lee, J.J.; Lötvall, J.O. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef]
  97. 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]
  98. Schneider, D.J.; Speth, J.M.; Penke, L.R.; Wettlaufer, S.H.; Swanson, J.A.; Peters-Golden, M. Mechanisms and modulation of microvesicle uptake in a model of alveolar cell communication. J. Biol. Chem. 2017, 292, 20897–20910. [Google Scholar] [CrossRef] [PubMed]
  99. Sharma, A. Role of stem cell derived exosomes in tumor biology. Int. J. Cancer 2018, 142, 1086–1092. [Google Scholar] [CrossRef]
  100. Kong, F.L.; Wang, X.P.; Li, Y.N.; Wang, H.X. The role of exosomes derived from cerebrospinal fluid of spinal cord injury in neuron proliferation in vitro. Artif. Cells Nanomed. Biotechnol. 2018, 46, 200–205. [Google Scholar] [CrossRef]
  101. Kooijmans, S.A.; Vader, P.; van Dommelen, S.M.; van Solinge, W.W.; Schiffelers, R.M. Exosome mimetics: A novel class of drug delivery systems. Int. J. Nanomed. 2012, 7, 1525–1541. [Google Scholar] [CrossRef]
  102. Wang, Y.; Zhang, M. Urinary Exosomes: A Promising Biomarker for Disease Diagnosis. Lab. Med. 2023, 54, 115–125. [Google Scholar] [CrossRef] [PubMed]
  103. Parizadeh, S.M.; Jafarzadeh-Esfehani, R.; Ghandehari, M.; Parizadeh, S.M.R.; Hassanian, S.M.; Rezayi, M.; Ghayour-Mobarhan, M.; Ferns, G.A.; Avan, A. Circulating Exosomes as Potential Biomarkers in Cardiovascular Disease. Curr. Pharm. Des. 2018, 24, 4436–4444. [Google Scholar] [CrossRef]
  104. Fan, Y.; Chen, Z.; Zhang, M. Role of exosomes in the pathogenesis, diagnosis, and treatment of central nervous system diseases. J. Transl. Med. 2022, 20, 291. [Google Scholar] [CrossRef] [PubMed]
  105. Shetgaonkar, G.G.; Marques, S.M.; CEM, D.C.; Vibhavari, R.J.A.; Kumar, L.; Shirodkar, R.K. Exosomes as cell-derivative carriers in the diagnosis and treatment of central nervous system diseases. Drug Deliv. Transl. Res. 2022, 12, 1047–1079. [Google Scholar] [CrossRef] [PubMed]
  106. Deng, Y.; Sun, Z.; Wang, L.; Wang, M.; Yang, J.; Li, G. Biosensor-based assay of exosome biomarker for early diagnosis of cancer. Front. Med. 2022, 16, 157–175. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, X.; Huang, J.; Chen, W.; Li, G.; Li, Z.; Lei, J. The updated role of exosomal proteins in the diagnosis, prognosis, and treatment of cancer. Exp. Mol. Med. 2022, 54, 1390–1400. [Google Scholar] [CrossRef] [PubMed]
  108. De Jong, O.G.; Verhaar, M.C.; Chen, Y.; Vader, P.; Gremmels, H.; Posthuma, G.; Schiffelers, R.M.; Gucek, M.; van Balkom, B.W. Cellular stress conditions are reflected in the protein and RNA content of endothelial cell-derived exosomes. J. Extracell. Vesicles 2012, 1, 18396. [Google Scholar] [CrossRef] [PubMed]
  109. Hergenreider, E.; Heydt, S.; Tréguer, K.; Boettger, T.; Horrevoets, A.J.; Zeiher, A.M.; Scheffer, M.P.; Frangakis, A.S.; Yin, X.; Mayr, M.; et al. Atheroprotective communication between endothelial cells and smooth muscle cells through miRNAs. Nat. Cell Biol. 2012, 14, 249–256. [Google Scholar] [CrossRef] [PubMed]
  110. Ailawadi, S.; Wang, X.; Gu, H.; Fan, G.C. Pathologic function and therapeutic potential of exosomes in cardiovascular disease. Biochim. Biophys. Acta 2015, 1852, 1–11. [Google Scholar] [CrossRef] [PubMed]
  111. Ghafarian, F.; Pashirzad, M.; Khazaei, M.; Rezayi, M.; Hassanian, S.M.; Ferns, G.A.; Avan, A. The clinical impact of exosomes in cardiovascular disorders: From basic science to clinical application. J. Cell. Physiol. 2019, 234, 12226–12236. [Google Scholar] [CrossRef]
  112. Zheng, D.; Huo, M.; Li, B.; Wang, W.; Piao, H.; Wang, Y.; Zhu, Z.; Li, D.; Wang, T.; Liu, K. The Role of Exosomes and Exosomal MicroRNA in Cardiovascular Disease. Front. Cell Dev. Biol. 2020, 8, 616161. [Google Scholar] [CrossRef]
  113. Kakkar, R.; Lee, R.T. Intramyocardial fibroblast myocyte communication. Circ. Res. 2010, 106, 47–57. [Google Scholar] [CrossRef] [PubMed]
  114. Lionetti, V.; Bianchi, G.; Recchia, F.A.; Ventura, C. Control of autocrine and paracrine myocardial signals: An emerging therapeutic strategy in heart failure. Heart Fail. Rev. 2010, 15, 531–542. [Google Scholar] [CrossRef]
  115. Li, X.X.; Yang, L.X.; Wang, C.; Li, H.; Shi, D.S.; Wang, J. The Roles of Exosomal Proteins: Classification, Function, and Applications. Int. J. Mol. Sci. 2023, 24, 3061. [Google Scholar] [CrossRef] [PubMed]
  116. Chinnappan, R.; Ramadan, Q.; Zourob, M. An integrated lab-on-a-chip platform for pre-concentration and detection of colorectal cancer exosomes using anti-CD63 aptamer as a recognition element. Biosens. Bioelectron. 2023, 220, 114856. [Google Scholar] [CrossRef]
  117. Li, L.; Wen, J.; Li, H.; He, Y.; Cui, X.; Zhang, X.; Guan, X.; Li, Z.; Cheng, M. Exosomal circ-1199 derived from EPCs exposed to oscillating shear stress acts as a sponge of let-7g-5p to promote endothelial-mesenchymal transition of EPCs by increasing HMGA2 expression. Life Sci. 2023, 312, 121223. [Google Scholar] [CrossRef] [PubMed]
  118. Hao, Y.; Liu, J.; Liu, J.; Yang, N.; Smith, S.C., Jr.; Huo, Y.; Fonarow, G.C.; Ge, J.; Taubert, K.A.; Morgan, L.; et al. Sex Differences in In-Hospital Management and Outcomes of Patients with Acute Coronary Syndrome. Circulation 2019, 139, 1776–1785. [Google Scholar] [CrossRef] [PubMed]
  119. Benjamin, E.J.; Muntner, P.; Alonso, A.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Chang, A.R.; Cheng, S.; Das, S.R.; et al. Heart Disease and Stroke Statistics-2019 Update: A Report From the American Heart Association. Circulation 2019, 139, e56–e528. [Google Scholar] [CrossRef] [PubMed]
  120. Neumann, F.J.; Sousa-Uva, M.; Ahlsson, A.; Alfonso, F.; Banning, A.P.; Benedetto, U.; Byrne, R.A.; Collet, J.P.; Falk, V.; Head, S.J.; et al. 2018 ESC/EACTS Guidelines on myocardial revascularization. Eur. Heart J. 2019, 40, 87–165. [Google Scholar] [CrossRef]
  121. Wu, G.; Zhang, J.; Zhao, Q.; Zhuang, W.; Ding, J.; Zhang, C.; Gao, H.; Pang, D.W.; Pu, K.; Xie, H.Y. Molecularly Engineered Macrophage-Derived Exosomes with Inflammation Tropism and Intrinsic Heme Biosynthesis for Atherosclerosis Treatment. Angew. Chem. 2020, 59, 4068–4074. [Google Scholar] [CrossRef] [PubMed]
  122. Deanfield, J.E.; Halcox, J.P.; Rabelink, T.J. Endothelial function and dysfunction: Testing and clinical relevance. Circulation 2007, 115, 1285–1295. [Google Scholar] [CrossRef] [PubMed]
  123. Mudau, M.; Genis, A.; Lochner, A.; Strijdom, H. Endothelial dysfunction: The early predictor of atherosclerosis. Cardiovasc. J. Afr. 2012, 23, 222–231. [Google Scholar] [CrossRef]
  124. Wadley, A.J.; Veldhuijzen van Zanten, J.J.; Aldred, S. The interactions of oxidative stress and inflammation with vascular dysfunction in ageing: The vascular health triad. Age 2013, 35, 705–718. [Google Scholar] [CrossRef] [PubMed]
  125. Sun, X.; Feinberg, M.W. Regulation of endothelial cell metabolism: Just go with the flow. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 13–15. [Google Scholar] [CrossRef]
  126. Libby, P.; Okamoto, Y.; Rocha, V.Z.; Folco, E. Inflammation in atherosclerosis: Transition from theory to practice. Circ. J. Off. J. Jpn. Circ. Soc. 2010, 74, 213–220. [Google Scholar] [CrossRef]
  127. Tang, N.; Sun, B.; Gupta, A.; Rempel, H.; Pulliam, L. Monocyte exosomes induce adhesion molecules and cytokines via activation of NF-κB in endothelial cells. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2016, 30, 3097–3106. [Google Scholar] [CrossRef]
  128. De Silva, N.; Samblas, M.; Martínez, J.A.; Milagro, F.I. Effects of exosomes from LPS-activated macrophages on adipocyte gene expression, differentiation, and insulin-dependent glucose uptake. J. Physiol. Biochem. 2018, 74, 559–568. [Google Scholar] [CrossRef]
  129. Zhan, R.; Leng, X.; Liu, X.; Wang, X.; Gong, J.; Yan, L.; Wang, L.; Wang, Y.; Wang, X.; Qian, L.J. Heat shock protein 70 is secreted from endothelial cells by a non-classical pathway involving exosomes. Biochem. Biophys. Res. Commun. 2009, 387, 229–233. [Google Scholar] [CrossRef]
  130. Selmaj, I.; Mycko, M.P.; Raine, C.S.; Selmaj, K.W. The role of exosomes in CNS inflammation and their involvement in multiple sclerosis. J. Neuroimmunol. 2017, 306, 1–10. [Google Scholar] [CrossRef]
  131. Poller, W.; Dimmeler, S.; Heymans, S.; Zeller, T.; Haas, J.; Karakas, M.; Leistner, D.M.; Jakob, P.; Nakagawa, S.; Blankenberg, S.; et al. Non-coding RNAs in cardiovascular diseases: Diagnostic and therapeutic perspectives. Eur. Heart J. 2018, 39, 2704–2716. [Google Scholar] [CrossRef]
  132. Zhang, J.; Li, S.; Li, L.; Li, M.; Guo, C.; Yao, J.; Mi, S. Exosome and exosomal microRNA: Trafficking, sorting, and function. Genom. Proteom. Bioinform. 2015, 13, 17–24. [Google Scholar] [CrossRef]
  133. Mittelbrunn, M.; Gutiérrez-Vázquez, C.; Villarroya-Beltri, C.; González, S.; Sánchez-Cabo, F.; González, M.; Bernad, A.; Sánchez-Madrid, F. Unidirectional transfer of microRNA-loaded exosomes from T cells to antigen-presenting cells. Nat. Commun. 2011, 2, 282. [Google Scholar] [CrossRef]
  134. Collino, F.; Deregibus, M.C.; Bruno, S.; Sterpone, L.; Aghemo, G.; Viltono, L.; Tetta, C.; Camussi, G. Microvesicles derived from adult human bone marrow and tissue specific mesenchymal stem cells shuttle selected pattern of miRNAs. PLoS ONE 2010, 5, e11803. [Google Scholar] [CrossRef] [PubMed]
  135. Eirin, A.; Zhu, X.Y.; Puranik, A.S.; Woollard, J.R.; Tang, H.; Dasari, S.; Lerman, A.; van Wijnen, A.J.; Lerman, L.O. Comparative proteomic analysis of extracellular vesicles isolated from porcine adipose tissue-derived mesenchymal stem/stromal cells. Sci. Rep. 2016, 6, 36120. [Google Scholar] [CrossRef] [PubMed]
  136. Hunter, M.P.; Ismail, N.; Zhang, X.; Aguda, B.D.; Lee, E.J.; Yu, L.; Xiao, T.; Schafer, J.; Lee, M.L.; Schmittgen, T.D.; et al. Detection of microRNA expression in human peripheral blood microvesicles. PLoS ONE 2008, 3, e3694. [Google Scholar] [CrossRef]
  137. Zhang, Y.; Hu, Y.W.; Zheng, L.; Wang, Q. Characteristics and Roles of Exosomes in Cardiovascular Disease. DNA Cell Biol. 2017, 36, 202–211. [Google Scholar] [CrossRef] [PubMed]
  138. Boulanger, C.M.; Loyer, X.; Rautou, P.E.; Amabile, N. Extracellular vesicles in coronary artery disease. Nat. Rev. Cardiol. 2017, 14, 259–272. [Google Scholar] [CrossRef]
  139. Van Rooij, E.; Olson, E.N. MicroRNA therapeutics for cardiovascular disease: Opportunities and obstacles. Nat. Rev. Drug Discov. 2012, 11, 860–872. [Google Scholar] [CrossRef]
  140. Njock, M.S.; Cheng, H.S.; Dang, L.T.; Nazari-Jahantigh, M.; Lau, A.C.; Boudreau, E.; Roufaiel, M.; Cybulsky, M.I.; Schober, A.; Fish, J.E. Endothelial cells suppress monocyte activation through secretion of extracellular vesicles containing antiinflammatory microRNAs. Blood 2015, 125, 3202–3212. [Google Scholar] [CrossRef]
  141. Wang, Z.; Zhang, J.; Zhang, S.; Yan, S.; Wang, Z.; Wang, C.; Zhang, X. MiR-30e and miR-92a are related to atherosclerosis by targeting ABCA1. Mol. Med. Rep. 2019, 19, 3298–3304. [Google Scholar] [CrossRef] [PubMed]
  142. Wang, Y.; Zhao, R.; Liu, W.; Wang, Z.; Rong, J.; Long, X.; Liu, Z.; Ge, J.; Shi, B. Exosomal circHIPK3 Released from Hypoxia-Pretreated Cardiomyocytes Regulates Oxidative Damage in Cardiac Microvascular Endothelial Cells via the miR-29a/IGF-1 Pathway. Oxidative Med. Cell. Longev. 2019, 2019, 7954657. [Google Scholar] [CrossRef] [PubMed]
  143. Linna-Kuosmanen, S.; Tomas Bosch, V.; Moreau, P.R.; Bouvy-Liivrand, M.; Niskanen, H.; Kansanen, E.; Kivelä, A.; Hartikainen, J.; Hippeläinen, M.; Kokki, H.; et al. NRF2 is a key regulator of endothelial microRNA expression under proatherogenic stimuli. Cardiovasc. Res. 2021, 117, 1339–1357. [Google Scholar] [CrossRef]
  144. Bi, S.; Wang, C.; Jin, Y.; Lv, Z.; Xing, X.; Lu, Q. Correlation between serum exosome derived miR-208a and acute coronary syndrome. Int. J. Clin. Exp. Med. 2015, 8, 4275–4280. [Google Scholar] [PubMed]
  145. Wang, Y.; Liang, J.; Xu, J.; Wang, X.; Zhang, X.; Wang, W.; Chen, L.; Yuan, T. Circulating exosomes and exosomal lncRNA HIF1A-AS1 in atherosclerosis. Int. J. Clin. Exp. Pathol. 2017, 10, 8383–8388. [Google Scholar]
  146. Emanueli, C.; Shearn, A.I.; Laftah, A.; Fiorentino, F.; Reeves, B.C.; Beltrami, C.; Mumford, A.; Clayton, A.; Gurney, M.; Shantikumar, S.; et al. Coronary Artery-Bypass-Graft Surgery Increases the Plasma Concentration of Exosomes Carrying a Cargo of Cardiac MicroRNAs: An Example of Exosome Trafficking Out of the Human Heart with Potential for Cardiac Biomarker Discovery. PLoS ONE 2016, 11, e0154274. [Google Scholar] [CrossRef]
  147. Fasolo, F.; Di Gregoli, K.; Maegdefessel, L.; Johnson, J.L. Non-coding RNAs in cardiovascular cell biology and atherosclerosis. Cardiovasc. Res. 2019, 115, 1732–1756. [Google Scholar] [CrossRef] [PubMed]
  148. Han, Z.; Chen, H.; Guo, Z.; Shen, J.; Luo, W.; Xie, F.; Wan, Y.; Wang, S.; Li, J.; He, J. Circular RNAs and Their Role in Exosomes. Front. Oncol. 2022, 12, 848341. [Google Scholar] [CrossRef]
  149. Wu, W.P.; Pan, Y.H.; Cai, M.Y.; Cen, J.M.; Chen, C.; Zheng, L.; Liu, X.; Xiong, X.D. Plasma-Derived Exosomal Circular RNA hsa_circ_0005540 as a Novel Diagnostic Biomarker for Coronary Artery Disease. Dis. Markers 2020, 2020, 3178642. [Google Scholar] [CrossRef]
  150. Shi, C.; Ulke-Lemée, A.; Deng, J.; Batulan, Z.; O’Brien, E.R. Characterization of heat shock protein 27 in extracellular vesicles: A potential anti-inflammatory therapy. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2019, 33, 1617–1630. [Google Scholar] [CrossRef]
  151. Liu, Y.; Li, C.; Wu, H.; Xie, X.; Sun, Y.; Dai, M. Paeonol Attenuated Inflammatory Response of Endothelial Cells via Stimulating Monocytes-Derived Exosomal MicroRNA-223. Front. Pharmacol. 2018, 9, 1105. [Google Scholar] [CrossRef]
  152. Feld, S.; Kjellgren, O.; Smalling, R.W. Aggressive interventional treatment of acute myocardial infarction. Lessons from the animal laboratory applied to the catheterization suite. Cardiology 1995, 86, 365–373. [Google Scholar] [CrossRef]
  153. Chen, X.; Huang, F.; Liu, Y.; Liu, S.; Tan, G. Exosomal miR-152-5p and miR-3681-5p function as potential biomarkers for ST-segment elevation myocardial infarction. Clinics 2022, 77, 100038. [Google Scholar] [CrossRef] [PubMed]
  154. Gong, F.F.; Vaitenas, I.; Malaisrie, S.C.; Maganti, K. Mechanical Complications of Acute Myocardial Infarction: A Review. JAMA Cardiol. 2021, 6, 341–349. [Google Scholar] [CrossRef]
  155. Doenst, T.; Haverich, A.; Serruys, P.; Bonow, R.O.; Kappetein, P.; Falk, V.; Velazquez, E.; Diegeler, A.; Sigusch, H. PCI and CABG for Treating Stable Coronary Artery Disease: JACC Review Topic of the Week. J. Am. Coll. Cardiol. 2019, 73, 964–976. [Google Scholar] [CrossRef] [PubMed]
  156. Sonny, A.; Joseph, L. Improving CABG Mortality Further: Striving Toward Perfection. J. Am. Coll. Cardiol. 2021, 78, 123–125. [Google Scholar] [CrossRef] [PubMed]
  157. Jenča, D.; Melenovský, V.; Stehlik, J.; Staněk, V.; Kettner, J.; Kautzner, J.; Adámková, V.; Wohlfahrt, P. Heart failure after myocardial infarction: Incidence and predictors. ESC Heart Fail. 2021, 8, 222–237. [Google Scholar] [CrossRef]
  158. Pamukcu, H.E.; Felekoğlu, M.A.; Algül, E.; Şahan, H.F.; Aydinyilmaz, F.; Guliyev, İ.; İnci, S.D.; Özbeyaz, N.B.; Nallbani, A. Copeptin levels predict left ventricular systolic function in STEMI patients: A 2D speckle tracking echocardiography-based prospective observational study. Medicine 2020, 99, e23514. [Google Scholar] [CrossRef]
  159. Van der Bijl, P.; Abou, R.; Goedemans, L.; Gersh, B.J.; Holmes, D.R., Jr.; Ajmone Marsan, N.; Delgado, V.; Bax, J.J. Left Ventricular Post-Infarct Remodeling: Implications for Systolic Function Improvement and Outcomes in the Modern Era. JACC. Heart Fail. 2020, 8, 131–140. [Google Scholar] [CrossRef]
  160. Vogel, B.; Claessen, B.E.; Arnold, S.V.; Chan, D.; Cohen, D.J.; Giannitsis, E.; Gibson, C.M.; Goto, S.; Katus, H.A.; Kerneis, M.; et al. ST-segment elevation myocardial infarction. Nat. Rev. Dis. Primers 2019, 5, 39. [Google Scholar] [CrossRef]
  161. Lechner, I.; Reindl, M.; Metzler, B.; Reinstadler, S.J. Predictors of Long-Term Outcome in STEMI and NSTEMI-Insights from J-MINUET. J. Clin. Med. 2020, 9, 3166. [Google Scholar] [CrossRef] [PubMed]
  162. Cohen, A.; Greco, J.; Levitus, M.; Nelson, M. The use of point-of-care ultrasound to diagnose infective endocarditis causing an NSTEMI in a patient with chest pain. J. Am. Coll. Emerg. Physicians Open 2020, 1, 120–123. [Google Scholar] [CrossRef]
  163. Saad, M.; Stiermaier, T.; Fuernau, G.; Pöss, J.; Desch, S.; Thiele, H.; Eitel, I. Impact of chronic total occlusion in a non-infarct-related coronary artery on myocardial injury assessed by cardiac magnetic resonance imaging and prognosis in ST-elevation myocardial infarction. Int. J. Cardiol. 2018, 265, 251–255. [Google Scholar] [CrossRef]
  164. Singla, D.K. Stem cells and exosomes in cardiac repair. Curr. Opin. Pharmacol. 2016, 27, 19–23. [Google Scholar] [CrossRef] [PubMed]
  165. Cheng, C.; Wang, Q.; You, W.; Chen, M.; Xia, J. MiRNAs as biomarkers of myocardial infarction: A meta-analysis. PLoS ONE 2014, 9, e88566. [Google Scholar] [CrossRef] [PubMed]
  166. Deddens, J.C.; Vrijsen, K.R.; Colijn, J.M.; Oerlemans, M.I.; Metz, C.H.; van der Vlist, E.J.; Nolte-’t Hoen, E.N.; den Ouden, K.; Jansen Of Lorkeers, S.J.; van der Spoel, T.I.; et al. Circulating Extracellular Vesicles Contain miRNAs and are Released as Early Biomarkers for Cardiac Injury. J. Cardiovasc. Transl. Res. 2016, 9, 291–301. [Google Scholar] [CrossRef]
  167. Zheng, M.L.; Liu, X.Y.; Han, R.J.; Yuan, W.; Sun, K.; Zhong, J.C.; Yang, X.C. Circulating exosomal long non-coding RNAs in patients with acute myocardial infarction. J. Cell. Mol. Med. 2020, 24, 9388–9396. [Google Scholar] [CrossRef]
  168. Chen, Z.; Yan, Y.; Wu, J.; Qi, C.; Liu, J.; Wang, J. Expression level and diagnostic value of exosomal NEAT1/miR-204/MMP-9 in acute ST-segment elevation myocardial infarction. IUBMB Life 2020, 72, 2499–2507. [Google Scholar] [CrossRef]
  169. Ibanez, B.; Rossello, X. Left Ventricular Remodeling Is No Longer a Relevant Outcome After Myocardial Infarction. JACC. Cardiovasc. Imaging 2019, 12, 2457–2459. [Google Scholar] [CrossRef]
  170. Dorn, G.W., 2nd. Novel pharmacotherapies to abrogate postinfarction ventricular remodeling. Nat. Rev. Cardiol. 2009, 6, 283–291. [Google Scholar] [CrossRef] [PubMed]
  171. Samouillan, V.; Martinez de Lejarza Samper, I.M.; Amaro, A.B.; Vilades, D.; Dandurand, J.; Casas, J.; Jorge, E.; de Gonzalo Calvo, D.; Gallardo, A.; Lerma, E.; et al. Biophysical and Lipidomic Biomarkers of Cardiac Remodeling Post-Myocardial Infarction in Humans. Biomolecules 2020, 10, 1471. [Google Scholar] [CrossRef] [PubMed]
  172. Silva-Palacios, A.; Arroyo-Campuzano, M.; Flores-García, M.; Patlán, M.; Hernández-Díazcouder, A.; Alcántara, D.; Ramírez-Camacho, I.; Arana-Hidalgo, D.; Soria-Castro, E.; Sánchez, F.; et al. Citicoline Modifies the Expression of Specific miRNAs Related to Cardioprotection in Patients with ST-Segment Elevation Myocardial Infarction Subjected to Coronary Angioplasty. Pharmaceuticals 2022, 15, 925. [Google Scholar] [CrossRef]
  173. Guan, R.; Zeng, K.; Zhang, B.; Gao, M.; Li, J.; Jiang, H.; Liu, Y.; Qiang, Y.; Liu, Z.; Li, J.; et al. Plasma Exosome miRNAs Profile in Patients With ST-Segment Elevation Myocardial Infarction. Front. Cardiovasc. Med. 2022, 9, 848812. [Google Scholar] [CrossRef]
  174. Vicencio, J.M.; Yellon, D.M.; Sivaraman, V.; Das, D.; Boi-Doku, C.; Arjun, S.; Zheng, Y.; Riquelme, J.A.; Kearney, J.; Sharma, V.; et al. Plasma exosomes protect the myocardium from ischemia-reperfusion injury. J. Am. Coll. Cardiol. 2015, 65, 1525–1536. [Google Scholar] [CrossRef] [PubMed]
  175. Aurora, A.B.; Mahmoud, A.I.; Luo, X.; Johnson, B.A.; van Rooij, E.; Matsuzaki, S.; Humphries, K.M.; Hill, J.A.; Bassel-Duby, R.; Sadek, H.A.; et al. MicroRNA-214 protects the mouse heart from ischemic injury by controlling Ca2+ overload and cell death. J. Clin. Investig. 2012, 122, 1222–1232. [Google Scholar] [CrossRef] [PubMed]
  176. Van Balkom, B.W.; de Jong, O.G.; Smits, M.; Brummelman, J.; den Ouden, K.; de Bree, P.M.; van Eijndhoven, M.A.; Pegtel, D.M.; Stoorvogel, W.; Würdinger, T.; et al. Endothelial cells require miR-214 to secrete exosomes that suppress senescence and induce angiogenesis in human and mouse endothelial cells. Blood 2013, 121, 3997–4006. [Google Scholar] [CrossRef] [PubMed]
  177. Wang, Y.; Zhao, R.; Shen, C.; Liu, W.; Yuan, J.; Li, C.; Deng, W.; Wang, Z.; Zhang, W.; Ge, J.; et al. Exosomal CircHIPK3 Released from Hypoxia-Induced Cardiomyocytes Regulates Cardiac Angiogenesis after Myocardial Infarction. Oxidative Med. Cell. Longev. 2020, 2020, 8418407. [Google Scholar] [CrossRef]
  178. Chen, X.; Luo, Q. Potential clinical applications of exosomes in the diagnosis, treatment, and prognosis of cardiovascular diseases: A narrative review. Ann. Transl. Med. 2022, 10, 372. [Google Scholar] [CrossRef]
  179. Gray, W.D.; French, K.M.; Ghosh-Choudhary, S.; Maxwell, J.T.; Brown, M.E.; Platt, M.O.; Searles, C.D.; Davis, M.E. Identification of therapeutic covariant microRNA clusters in hypoxia-treated cardiac progenitor cell exosomes using systems biology. Circ. Res. 2015, 116, 255–263. [Google Scholar] [CrossRef]
  180. Kervadec, A.; Bellamy, V.; El Harane, N.; Arakélian, L.; Vanneaux, V.; Cacciapuoti, I.; Nemetalla, H.; Périer, M.C.; Toeg, H.D.; Richart, A.; et al. Cardiovascular progenitor-derived extracellular vesicles recapitulate the beneficial effects of their parent cells in the treatment of chronic heart failure. J. Heart Lung Transplant. Off. Publ. Int. Soc. Heart Transplant. 2016, 35, 795–807. [Google Scholar] [CrossRef]
  181. Barile, L.; Lionetti, V.; Cervio, E.; Matteucci, M.; Gherghiceanu, M.; Popescu, L.M.; Torre, T.; Siclari, F.; Moccetti, T.; Vassalli, G. Extracellular vesicles from human cardiac progenitor cells inhibit cardiomyocyte apoptosis and improve cardiac function after myocardial infarction. Cardiovasc. Res. 2014, 103, 530–541. [Google Scholar] [CrossRef] [PubMed]
  182. Zheng, J.; Zhang, X.; Cai, W.; Yang, Y.; Guo, T.; Li, J.; Dai, H. Bone Marrow Mesenchymal Stem Cell-Derived Exosomal microRNA-29b-3p Promotes Angiogenesis and Ventricular Remodeling in Rats with Myocardial Infarction by Targeting ADAMTS16. Cardiovasc. Toxicol. 2022, 22, 689–700. [Google Scholar] [CrossRef] [PubMed]
  183. Peng, Y.; Zhao, J.L.; Peng, Z.Y.; Xu, W.F.; Yu, G.L. Exosomal miR-25-3p from mesenchymal stem cells alleviates myocardial infarction by targeting pro-apoptotic proteins and EZH2. Cell Death Dis. 2020, 11, 317. [Google Scholar] [CrossRef]
  184. Lai, R.C.; Arslan, F.; Lee, M.M.; Sze, N.S.; Choo, A.; Chen, T.S.; Salto-Tellez, M.; Timmers, L.; Lee, C.N.; El Oakley, R.M.; et al. Exosome secreted by MSC reduces myocardial ischemia/reperfusion injury. Stem Cell Res. 2010, 4, 214–222. [Google Scholar] [CrossRef]
  185. Arslan, F.; Lai, R.C.; Smeets, M.B.; Akeroyd, L.; Choo, A.; Aguor, E.N.; Timmers, L.; van Rijen, H.V.; Doevendans, P.A.; Pasterkamp, G.; et al. Mesenchymal stem cell-derived exosomes increase ATP levels, decrease oxidative stress and activate PI3K/Akt pathway to enhance myocardial viability and prevent adverse remodeling after myocardial ischemia/reperfusion injury. Stem Cell Res. 2013, 10, 301–312. [Google Scholar] [CrossRef] [PubMed]
  186. Dharmarajan, K.; Rich, M.W. Epidemiology, Pathophysiology, and Prognosis of Heart Failure in Older Adults. Heart Fail. Clin. 2017, 13, 417–426. [Google Scholar] [CrossRef]
  187. Xue, R.; Tan, W.; Wu, Y.; Dong, B.; Xie, Z.; Huang, P.; He, J.; Dong, Y.; Liu, C. Role of Exosomal miRNAs in Heart Failure. Front. Cardiovasc. Med. 2020, 7, 592412. [Google Scholar] [CrossRef]
  188. Gao, Y.; Gao, Y.; Zhu, R.; Tan, X. Shenfu injection combined with furosemide in the treatment of chronic heart failure in patients with coronary heart disease: A protocol of randomized controlled trial. Medicine 2021, 100, e24113. [Google Scholar] [CrossRef] [PubMed]
  189. Sánchez Marteles, M.; Urrutia, A. [Acute heart failure: Acute cardiogenic pulmonary edema and cardiogenic shock]. Med. Clin. 2014, 142 (Suppl. S1), 14–19. [Google Scholar] [CrossRef]
  190. Su, M.; Wang, J.; Wang, C.; Wang, X.; Dong, W.; Qiu, W.; Wang, Y.; Zhao, X.; Zou, Y.; Song, L.; et al. MicroRNA-221 inhibits autophagy and promotes heart failure by modulating the p27/CDK2/mTOR axis. Cell Death Differ. 2015, 22, 986–999. [Google Scholar] [CrossRef] [PubMed]
  191. Kuwabara, Y.; Ono, K.; Horie, T.; Nishi, H.; Nagao, K.; Kinoshita, M.; Watanabe, S.; Baba, O.; Kojima, Y.; Shizuta, S.; et al. Increased microRNA-1 and microRNA-133a levels in serum of patients with cardiovascular disease indicate myocardial damage. Circulation. Cardiovasc. Genet. 2011, 4, 446–454. [Google Scholar] [CrossRef]
  192. LaFramboise, W.A.; Scalise, D.; Stoodley, P.; Graner, S.R.; Guthrie, R.D.; Magovern, J.A.; Becich, M.J. Cardiac fibroblasts influence cardiomyocyte phenotype in vitro. Am. J. Physiol. Cell Physiol. 2007, 292, C1799–C1808. [Google Scholar] [CrossRef]
  193. Fredj, S.; Bescond, J.; Louault, C.; Potreau, D. Interactions between cardiac cells enhance cardiomyocyte hypertrophy and increase fibroblast proliferation. J. Cell. Physiol. 2005, 202, 891–899. [Google Scholar] [CrossRef]
  194. Wang, B.; Wang, Z.M.; Ji, J.L.; Gan, W.; Zhang, A.; Shi, H.J.; Wang, H.; Lv, L.; Li, Z.; Tang, T.; et al. Macrophage-Derived Exosomal Mir-155 Regulating Cardiomyocyte Pyroptosis and Hypertrophy in Uremic Cardiomyopathy. JACC. Basic Transl. Sci. 2020, 5, 148–166. [Google Scholar] [CrossRef]
  195. Fang, X.; Stroud, M.J.; Ouyang, K.; Fang, L.; Zhang, J.; Dalton, N.D.; Gu, Y.; Wu, T.; Peterson, K.L.; Huang, H.D.; et al. Adipocyte-specific loss of PPARγ attenuates cardiac hypertrophy. JCI Insight 2016, 1, e89908. [Google Scholar] [CrossRef] [PubMed]
  196. Zou, M.; Wang, F.; Gao, R.; Wu, J.; Ou, Y.; Chen, X.; Wang, T.; Zhou, X.; Zhu, W.; Li, P.; et al. Autophagy inhibition of hsa-miR-19a-3p/19b-3p by targeting TGF-β R II during TGF-β1-induced fibrogenesis in human cardiac fibroblasts. Sci. Rep. 2016, 6, 24747. [Google Scholar] [CrossRef] [PubMed]
  197. Sun, L.Y.; Zhao, J.C.; Ge, X.M.; Zhang, H.; Wang, C.M.; Bie, Z.D. Circ_LAS1L regulates cardiac fibroblast activation, growth, and migration through miR-125b/SFRP5 pathway. Cell Biochem. Funct. 2020, 38, 443–450. [Google Scholar] [CrossRef] [PubMed]
  198. Yang, J.; Yu, X.; Xue, F.; Li, Y.; Liu, W.; Zhang, S. Exosomes derived from cardiomyocytes promote cardiac fibrosis via myocyte-fibroblast cross-talk. Am. J. Transl. Res. 2018, 10, 4350–4366. [Google Scholar]
  199. Liu, L.; Luo, F.; Lei, K. Exosomes Containing LINC00636 Inhibit MAPK1 through the miR-450a-2-3p Overexpression in Human Pericardial Fluid and Improve Cardiac Fibrosis in Patients with Atrial Fibrillation. Mediat. Inflamm. 2021, 2021, 9960241. [Google Scholar] [CrossRef]
  200. Yamaguchi, T.; Izumi, Y.; Nakamura, Y.; Yamazaki, T.; Shiota, M.; Sano, S.; Tanaka, M.; Osada-Oka, M.; Shimada, K.; Miura, K.; et al. Repeated remote ischemic conditioning attenuates left ventricular remodeling via exosome-mediated intercellular communication on chronic heart failure after myocardial infarction. Int. J. Cardiol. 2015, 178, 239–246. [Google Scholar] [CrossRef]
  201. Khan, M.; Nickoloff, E.; Abramova, T.; Johnson, J.; Verma, S.K.; Krishnamurthy, P.; Mackie, A.R.; Vaughan, E.; Garikipati, V.N.; Benedict, C.; et al. Embryonic stem cell-derived exosomes promote endogenous repair mechanisms and enhance cardiac function following myocardial infarction. Circ. Res. 2015, 117, 52–64. [Google Scholar] [CrossRef] [PubMed]
  202. Wang, C.; Zhang, C.; Liu, L.; A, X.; Chen, B.; Li, Y.; Du, J. Macrophage-Derived mir-155-Containing Exosomes Suppress Fibroblast Proliferation and Promote Fibroblast Inflammation during Cardiac Injury. Mol. Ther. J. Am. Soc. Gene Ther. 2017, 25, 192–204. [Google Scholar] [CrossRef]
  203. Gogiraju, R.; Bochenek, M.L.; Schäfer, K. Angiogenic Endothelial Cell Signaling in Cardiac Hypertrophy and Heart Failure. Front. Cardiovasc. Med. 2019, 6, 20. [Google Scholar] [CrossRef] [PubMed]
  204. Zhabyeyev, P.; Gandhi, M.; Mori, J.; Basu, R.; Kassiri, Z.; Clanachan, A.; Lopaschuk, G.D.; Oudit, G.Y. Pressure-overload-induced heart failure induces a selective reduction in glucose oxidation at physiological afterload. Cardiovasc. Res. 2013, 97, 676–685. [Google Scholar] [CrossRef]
  205. Qiao, L.; Hu, S.; Liu, S.; Zhang, H.; Ma, H.; Huang, K.; Li, Z.; Su, T.; Vandergriff, A.; Tang, J.; et al. microRNA-21-5p dysregulation in exosomes derived from heart failure patients impairs regenerative potential. J. Clin. Investig. 2019, 129, 2237–2250. [Google Scholar] [CrossRef] [PubMed]
  206. Liu, S.; Chen, J.; Shi, J.; Zhou, W.; Wang, L.; Fang, W.; Zhong, Y.; Chen, X.; Chen, Y.; Sabri, A.; et al. M1-like macrophage-derived exosomes suppress angiogenesis and exacerbate cardiac dysfunction in a myocardial infarction microenvironment. Basic Res. Cardiol. 2020, 115, 22. [Google Scholar] [CrossRef]
  207. Halkein, J.; Tabruyn, S.P.; Ricke-Hoch, M.; Haghikia, A.; Nguyen, N.Q.; Scherr, M.; Castermans, K.; Malvaux, L.; Lambert, V.; Thiry, M.; et al. MicroRNA-146a is a therapeutic target and biomarker for peripartum cardiomyopathy. J. Clin. Investig. 2013, 123, 2143–2154. [Google Scholar] [CrossRef]
  208. Ibrahim, A.G.; Cheng, K.; Marbán, E. Exosomes as critical agents of cardiac regeneration triggered by cell therapy. Stem Cell Rep. 2014, 2, 606–619. [Google Scholar] [CrossRef]
  209. Ribeiro-Rodrigues, T.M.; Laundos, T.L.; Pereira-Carvalho, R.; Batista-Almeida, D.; Pereira, R.; Coelho-Santos, V.; Silva, A.P.; Fernandes, R.; Zuzarte, M.; Enguita, F.J.; et al. Exosomes secreted by cardiomyocytes subjected to ischaemia promote cardiac angiogenesis. Cardiovasc. Res. 2017, 113, 1338–1350. [Google Scholar] [CrossRef] [PubMed]
  210. Lever, A.; Mackenzie, I. Sepsis: Definition, epidemiology, and diagnosis. BMJ 2007, 335, 879–883. [Google Scholar] [CrossRef]
  211. Russell, J.A.; Walley, K.R. Update in sepsis 2012. Am. J. Respir. Crit. Care Med. 2013, 187, 1303–1307. [Google Scholar] [CrossRef]
  212. Iskander, K.N.; Osuchowski, M.F.; Stearns-Kurosawa, D.J.; Kurosawa, S.; Stepien, D.; Valentine, C.; Remick, D.G. Sepsis: Multiple abnormalities, heterogeneous responses, and evolving understanding. Physiol. Rev. 2013, 93, 1247–1288. [Google Scholar] [CrossRef]
  213. Romero-Bermejo, F.J.; Ruiz-Bailen, M.; Gil-Cebrian, J.; Huertos-Ranchal, M.J. Sepsis-induced cardiomyopathy. Curr. Cardiol. Rev. 2011, 7, 163–183. [Google Scholar] [CrossRef] [PubMed]
  214. Blanco, J.; Muriel-Bombín, A.; Sagredo, V.; Taboada, F.; Gandía, F.; Tamayo, L.; Collado, J.; García-Labattut, A.; Carriedo, D.; Valledor, M.; et al. Incidence, organ dysfunction and mortality in severe sepsis: A Spanish multicentre study. Crit. Care 2008, 12, R158. [Google Scholar] [CrossRef] [PubMed]
  215. Parrillo, J.E.; Parker, M.M.; Natanson, C.; Suffredini, A.F.; Danner, R.L.; Cunnion, R.E.; Ognibene, F.P. Septic shock in humans. Advances in the understanding of pathogenesis, cardiovascular dysfunction, and therapy. Ann. Intern. Med. 1990, 113, 227–242. [Google Scholar] [CrossRef] [PubMed]
  216. Azevedo, L.C.; Janiszewski, M.; Pontieri, V.; Pedro Mde, A.; Bassi, E.; Tucci, P.J.; Laurindo, F.R. Platelet-derived exosomes from septic shock patients induce myocardial dysfunction. Crit. Care 2007, 11, R120. [Google Scholar] [CrossRef]
  217. Court, O.; Kumar, A.; Parrillo, J.E.; Kumar, A. Clinical review: Myocardial depression in sepsis and septic shock. Crit. Care 2002, 6, 500–508. [Google Scholar] [CrossRef]
  218. Riedemann, N.C.; Guo, R.F.; Ward, P.A. Novel strategies for the treatment of sepsis. Nat. Med. 2003, 9, 517–524. [Google Scholar] [CrossRef] [PubMed]
  219. Prabhu, S.D. Cytokine-induced modulation of cardiac function. Circ. Res. 2004, 95, 1140–1153. [Google Scholar] [CrossRef]
  220. Park, E.J.; Shimaoka, M.; Kiyono, H. Functional Flexibility of Exosomes and MicroRNAs of Intestinal Epithelial Cells in Affecting Inflammation. Front. Mol. Biosci. 2022, 9, 854487. [Google Scholar] [CrossRef] [PubMed]
  221. Tu, G.W.; Ma, J.F.; Li, J.K.; Su, Y.; Luo, J.C.; Hao, G.W.; Luo, M.H.; Cao, Y.R.; Zhang, Y.; Luo, Z. Exosome-Derived From Sepsis Patients’ Blood Promoted Pyroptosis of Cardiomyocytes by Regulating miR-885-5p/HMBOX1. Front. Cardiovasc. Med. 2022, 9, 774193. [Google Scholar] [CrossRef]
  222. Kosaka, N.; Iguchi, H.; Yoshioka, Y.; Takeshita, F.; Matsuki, Y.; Ochiya, T. Secretory mechanisms and intercellular transfer of microRNAs in living cells. J. Biol. Chem. 2010, 285, 17442–17452. [Google Scholar] [CrossRef] [PubMed]
  223. Kulshreshtha, A.; Ahmad, T.; Agrawal, A.; Ghosh, B. Proinflammatory role of epithelial cell-derived exosomes in allergic airway inflammation. J. Allergy Clin. Immunol. 2013, 131, 1194–1203.e14. [Google Scholar] [CrossRef] [PubMed]
  224. Li, J.; Liu, K.; Liu, Y.; Xu, Y.; Zhang, F.; Yang, H.; Liu, J.; Pan, T.; Chen, J.; Wu, M.; et al. Exosomes mediate the cell-to-cell transmission of IFN-α-induced antiviral activity. Nat. Immunol. 2013, 14, 793–803. [Google Scholar] [CrossRef]
  225. Essandoh, K.; Yang, L.; Wang, X.; Huang, W.; Qin, D.; Hao, J.; Wang, Y.; Zingarelli, B.; Peng, T.; Fan, G.C. Blockade of exosome generation with GW4869 dampens the sepsis-induced inflammation and cardiac dysfunction. Biochim. Biophys. Acta 2015, 1852, 2362–2371. [Google Scholar] [CrossRef]
  226. Janiszewski, M.; Do Carmo, A.O.; Pedro, M.A.; Silva, E.; Knobel, E.; Laurindo, F.R. Platelet-derived exosomes of septic individuals possess proapoptotic NAD(P)H oxidase activity: A novel vascular redox pathway. Crit. Care Med. 2004, 32, 818–825. [Google Scholar] [CrossRef]
  227. Day, B.J.; Fridovich, I.; Crapo, J.D. Manganic porphyrins possess catalase activity and protect endothelial cells against hydrogen peroxide-mediated injury. Arch. Biochem. Biophys. 1997, 347, 256–262. [Google Scholar] [CrossRef] [PubMed]
  228. Liu, D.; Shan, Y.; Valluru, L.; Bao, F. Mn (III) tetrakis (4-benzoic acid) porphyrin scavenges reactive species, reduces oxidative stress, and improves functional recovery after experimental spinal cord injury in rats: Comparison with methylprednisolone. BMC Neurosci. 2013, 14, 23. [Google Scholar] [CrossRef] [PubMed]
  229. Puhm, F.; Afonyushkin, T.; Resch, U.; Obermayer, G.; Rohde, M.; Penz, T.; Schuster, M.; Wagner, G.; Rendeiro, A.F.; Melki, I.; et al. Mitochondria Are a Subset of Extracellular Vesicles Released by Activated Monocytes and Induce Type I IFN and TNF Responses in Endothelial Cells. Circ. Res. 2019, 125, 43–52. [Google Scholar] [CrossRef]
  230. Raslova, K. An update on the treatment of type 1 and type 2 diabetes mellitus: Focus on insulin detemir, a long-acting human insulin analog. Vasc. Health Risk Manag. 2010, 6, 399–410. [Google Scholar] [CrossRef]
  231. Nakagami, H.; Kaneda, Y.; Ogihara, T.; Morishita, R. Endothelial dysfunction in hyperglycemia as a trigger of atherosclerosis. Curr. Diabetes Rev. 2005, 1, 59–63. [Google Scholar] [CrossRef] [PubMed]
  232. Cohen, G.; Riahi, Y.; Alpert, E.; Gruzman, A.; Sasson, S. The roles of hyperglycaemia and oxidative stress in the rise and collapse of the natural protective mechanism against vascular endothelial cell dysfunction in diabetes. Arch. Physiol. Biochem. 2007, 113, 259–267. [Google Scholar] [CrossRef]
  233. Joshi, M.; Kotha, S.R.; Malireddy, S.; Selvaraju, V.; Satoskar, A.R.; Palesty, A.; McFadden, D.W.; Parinandi, N.L.; Maulik, N. Conundrum of pathogenesis of diabetic cardiomyopathy: Role of vascular endothelial dysfunction, reactive oxygen species, and mitochondria. Mol. Cell. Biochem. 2014, 386, 233–249. [Google Scholar] [CrossRef]
  234. Wan, A.; Rodrigues, B. Endothelial cell-cardiomyocyte crosstalk in diabetic cardiomyopathy. Cardiovasc. Res. 2016, 111, 172–183. [Google Scholar] [CrossRef] [PubMed]
  235. 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]
  236. Huang-Doran, I.; Zhang, C.Y.; Vidal-Puig, A. Extracellular Vesicles: Novel Mediators of Cell Communication In Metabolic Disease. Trends Endocrinol. Metab. TEM 2017, 28, 3–18. [Google Scholar] [CrossRef] [PubMed]
  237. Salem, E.S.B.; Fan, G.C. Pathological Effects of Exosomes in Mediating Diabetic Cardiomyopathy. Adv. Exp. Med. Biol. 2017, 998, 113–138. [Google Scholar] [CrossRef]
  238. 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]
  239. Davidson, S.M.; Riquelme, J.A.; Takov, K.; Vicencio, J.M.; Boi-Doku, C.; Khoo, V.; Doreth, C.; Radenkovic, D.; Lavandero, S.; Yellon, D.M. Cardioprotection mediated by exosomes is impaired in the setting of type II diabetes but can be rescued by the use of non-diabetic exosomes in vitro. J. Cell. Mol. Med. 2018, 22, 141–151. [Google Scholar] [CrossRef]
  240. Yang, Y.; Wu, Y.; Hou, L.; Ge, X.; Song, G.; Jin, H. Heat shock protein 20 suppresses breast carcinogenesis by inhibiting the MAPK and AKT signaling pathways. Oncol. Lett. 2022, 24, 462. [Google Scholar] [CrossRef]
  241. 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]
  242. Tao, L.; Shi, J.; Yang, X.; Yang, L.; Hua, F. The Exosome: A New Player in Diabetic Cardiomyopathy. J. Cardiovasc. Transl. Res. 2019, 12, 62–67. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The three major categories of EVs: microvesicles, apoptotic bodies, and exosomes. Microvesicles are released through plasma membrane budding and have a size range of 50 nm to 1000 nm. Apoptotic bodies are released by cells undergoing apoptosis as blebs into the extracellular space and have diameter ranging from 50–5000 nm. Exosomes originate in the endosomal pathway through the formation of early-sorting endosomes, late-sorting endosomes, and eventually multivesicular bodies. They are released when multivesicular bodies fuse with the plasma membrane and range in size from 30 to 150 nm. This figure was created with BioRender.com.
Figure 1. The three major categories of EVs: microvesicles, apoptotic bodies, and exosomes. Microvesicles are released through plasma membrane budding and have a size range of 50 nm to 1000 nm. Apoptotic bodies are released by cells undergoing apoptosis as blebs into the extracellular space and have diameter ranging from 50–5000 nm. Exosomes originate in the endosomal pathway through the formation of early-sorting endosomes, late-sorting endosomes, and eventually multivesicular bodies. They are released when multivesicular bodies fuse with the plasma membrane and range in size from 30 to 150 nm. This figure was created with BioRender.com.
Ijms 24 15677 g001
Figure 2. Biogenesis, secretion, and identification of exosomes. Fluid and extracellular components, such as proteins, lipids, metabolites, small molecules, and ions, can enter cells through endocytosis and plasma membrane invagination, along with cell surface proteins. The plasma membrane bud formed on the side of the cell cavity shows outward-to-inward orientation of the plasma membrane. This budding process leads to the formation of early-sorting endosomes or possible fusion of the bud with early-sorting endosomes preformed by the trans-Golgi network (TGN) and mitochondrial-derived vesicles. The second invagination of late-sorting endosomes leads to the production of intraluminal vesicles, further modifying the cargo of future exosomes and allowing the entry of cytoplasmic components into the newly formed intraluminal vesicles. As part of intraluminal vesicles formation, proteins originally present on the cell surface can be clearly distributed within the intraluminal vesicles. Depending on the extent of invagination, this process can produce intraluminal vesicles of varying sizes and contents. Late-sorting endosomes produce multivesicular bodies with a well-defined set of intraluminal vesicles (future exosomes). Multivesicular bodies can either fuse with lysosomes for degradation or, if following a different trajectory, be transported to the plasma membrane via the cytoskeleton and microtubule networks. Exocytosis results in the release of exosomes, which possess lipid bilayers similar to the plasma membrane. Exosome biogenesis is associated with factors such as Ca2+, pH, SIRT1, and proteins like Rab GTPases and ESCRT, as discussed in the text. Additionally, several proteins are commonly used as exosome markers, including CD9, CD81, CD63, flotillin, TSG101, ceramides, and Alix. Exosome surface proteins encompass tetraspanins, integrins, membrane transport proteins, and more. Exosomes contain various cell surface proteins, intracellular proteins, RNA, and DNA. This figure was created with BioRender.com.
Figure 2. Biogenesis, secretion, and identification of exosomes. Fluid and extracellular components, such as proteins, lipids, metabolites, small molecules, and ions, can enter cells through endocytosis and plasma membrane invagination, along with cell surface proteins. The plasma membrane bud formed on the side of the cell cavity shows outward-to-inward orientation of the plasma membrane. This budding process leads to the formation of early-sorting endosomes or possible fusion of the bud with early-sorting endosomes preformed by the trans-Golgi network (TGN) and mitochondrial-derived vesicles. The second invagination of late-sorting endosomes leads to the production of intraluminal vesicles, further modifying the cargo of future exosomes and allowing the entry of cytoplasmic components into the newly formed intraluminal vesicles. As part of intraluminal vesicles formation, proteins originally present on the cell surface can be clearly distributed within the intraluminal vesicles. Depending on the extent of invagination, this process can produce intraluminal vesicles of varying sizes and contents. Late-sorting endosomes produce multivesicular bodies with a well-defined set of intraluminal vesicles (future exosomes). Multivesicular bodies can either fuse with lysosomes for degradation or, if following a different trajectory, be transported to the plasma membrane via the cytoskeleton and microtubule networks. Exocytosis results in the release of exosomes, which possess lipid bilayers similar to the plasma membrane. Exosome biogenesis is associated with factors such as Ca2+, pH, SIRT1, and proteins like Rab GTPases and ESCRT, as discussed in the text. Additionally, several proteins are commonly used as exosome markers, including CD9, CD81, CD63, flotillin, TSG101, ceramides, and Alix. Exosome surface proteins encompass tetraspanins, integrins, membrane transport proteins, and more. Exosomes contain various cell surface proteins, intracellular proteins, RNA, and DNA. This figure was created with BioRender.com.
Ijms 24 15677 g002
Figure 3. Biogenesis and secretion of exosomes and their contrasting biological functions in the cardiovascular system. The biogenesis and secretion of exosomes can play either a protective role (in diagnosis, as therapeutic vector, and in prognostic evaluations) or a harmful role (by inducing oxidative stress, endothelial dysfunction, and apoptosis) in the context of the cardiovascular system.
Figure 3. Biogenesis and secretion of exosomes and their contrasting biological functions in the cardiovascular system. The biogenesis and secretion of exosomes can play either a protective role (in diagnosis, as therapeutic vector, and in prognostic evaluations) or a harmful role (by inducing oxidative stress, endothelial dysfunction, and apoptosis) in the context of the cardiovascular system.
Ijms 24 15677 g003
Figure 4. Different functions of exosomal NcRNAs in CVD. Exosomal NcRNAs can have protective effects (blue) or harmful effects (red) in: (A) coronary artery disease, especially in the context of in coronary atherosclerosis, (B) myocardial infarction, encompassing both ST-segment elevation MI (TEMI) and non-ST-segment elevation MI (NSTEMI), and (C) heart failure scenarios such as cardiac hypertrophy, cardiac fibrosis, and myocardial angiogenesis. Addiontially, these exosomal NcRNAs can also serve as biomarkers (black). This figure was created with BioRender.com.
Figure 4. Different functions of exosomal NcRNAs in CVD. Exosomal NcRNAs can have protective effects (blue) or harmful effects (red) in: (A) coronary artery disease, especially in the context of in coronary atherosclerosis, (B) myocardial infarction, encompassing both ST-segment elevation MI (TEMI) and non-ST-segment elevation MI (NSTEMI), and (C) heart failure scenarios such as cardiac hypertrophy, cardiac fibrosis, and myocardial angiogenesis. Addiontially, these exosomal NcRNAs can also serve as biomarkers (black). This figure was created with BioRender.com.
Ijms 24 15677 g004
Figure 5. Overview of exosomes in other cardiomyopathy and the underlying mechanisms. Exosomes from cardiomyocytes of sepsis patients contain that can be transmitted to endothelial cells, contributing to the development of septic cardiomyopathy. Promising therapeutic strategies aimed at either inhibiting exosome production (e.g., using GW4869) or eliminating ROS within exosomes (by utilizing MnTBAP) may hold significant potential for patients with septic shock (for detailed information, please refer to the main text). Exosomes released by both diabetic and healthy cardiomyocytes, or those subjected to beneficial modifications under diabetic conditions, possess the capacity to either promote or inhibit the progression of diabetic cardiomyopathy. This figure was created with BioRender.com.
Figure 5. Overview of exosomes in other cardiomyopathy and the underlying mechanisms. Exosomes from cardiomyocytes of sepsis patients contain that can be transmitted to endothelial cells, contributing to the development of septic cardiomyopathy. Promising therapeutic strategies aimed at either inhibiting exosome production (e.g., using GW4869) or eliminating ROS within exosomes (by utilizing MnTBAP) may hold significant potential for patients with septic shock (for detailed information, please refer to the main text). Exosomes released by both diabetic and healthy cardiomyocytes, or those subjected to beneficial modifications under diabetic conditions, possess the capacity to either promote or inhibit the progression of diabetic cardiomyopathy. This figure was created with BioRender.com.
Ijms 24 15677 g005
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

Zhang, S.; Yang, Y.; Lv, X.; Liu, W.; Zhu, S.; Wang, Y.; Xu, H. Unraveling the Intricate Roles of Exosomes in Cardiovascular Diseases: A Comprehensive Review of Physiological Significance and Pathological Implications. Int. J. Mol. Sci. 2023, 24, 15677. https://doi.org/10.3390/ijms242115677

AMA Style

Zhang S, Yang Y, Lv X, Liu W, Zhu S, Wang Y, Xu H. Unraveling the Intricate Roles of Exosomes in Cardiovascular Diseases: A Comprehensive Review of Physiological Significance and Pathological Implications. International Journal of Molecular Sciences. 2023; 24(21):15677. https://doi.org/10.3390/ijms242115677

Chicago/Turabian Style

Zhang, Shuai, Yu Yang, Xinchen Lv, Wendong Liu, Shaohua Zhu, Ying Wang, and Hongfei Xu. 2023. "Unraveling the Intricate Roles of Exosomes in Cardiovascular Diseases: A Comprehensive Review of Physiological Significance and Pathological Implications" International Journal of Molecular Sciences 24, no. 21: 15677. https://doi.org/10.3390/ijms242115677

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