Next Article in Journal
Restructuring of Lamina-Associated Domains in Senescence and Cancer
Next Article in Special Issue
Large-Scale Profiling on lncRNAs in Human Platelets: Correlation with Platelet Reactivity
Previous Article in Journal
Differential Expression of CD45RO and CD45RA in Bovine T Cells
Previous Article in Special Issue
Group B Streptococcal Hemolytic Pigment Impairs Platelet Function in a Two-Step Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extracellular Vesicles as Drivers of Immunoinflammation in Atherothrombosis

1
Cardiovascular ICCC Program, Research Institute Hospital de la Santa Creu i Sant Pau, IIB-Sant Pau, 08025 Barcelona, Spain
2
Department of Pharmacological and Biomolecular Sciences, Università degli Studi di Milano, 20133 Milan, Italy
3
Centro de Investigación Biomédica en Red Cardiovascular (CIBERCV), Instituto de Salud Carlos III, 28029 Madrid, Spain
*
Author to whom correspondence should be addressed.
Cells 2022, 11(11), 1845; https://doi.org/10.3390/cells11111845
Submission received: 10 May 2022 / Revised: 30 May 2022 / Accepted: 1 June 2022 / Published: 5 June 2022

Abstract

:
Atherosclerotic cardiovascular disease is the leading cause of morbidity and mortality all over the world. Extracellular vesicles (EVs), small lipid-bilayer membrane vesicles released by most cellular types, exert pivotal and multifaceted roles in physiology and disease. Emerging evidence emphasizes the importance of EVs in intercellular communication processes with key effects on cell survival, endothelial homeostasis, inflammation, neoangiogenesis, and thrombosis. This review focuses on EVs as effective signaling molecules able to both derail vascular homeostasis and induce vascular dysfunction, inflammation, plaque progression, and thrombus formation as well as drive anti-inflammation, vascular repair, and atheroprotection. We provide a comprehensive and updated summary of the role of EVs in the development or regression of atherosclerotic lesions, highlighting the link between thrombosis and inflammation. Importantly, we also critically describe their potential clinical use as disease biomarkers or therapeutic agents in atherothrombosis.

1. Introduction

From the initial seminal discoveries [1,2] to the current state of research, the understanding of extracellular vesicles (EVs) has massively progressed and further advancements in the area are expected [3]. EVs are small lipid-bilayer vesicles that contain a complex cargo of nucleic acids, proteins, and lipids [4]. They are considered cell-to-cell signaling modules released from any cell type (both prokaryotic and eukaryotic) into distinct interstitial compartments and body fluids with potential value as disease biomarkers and drug-delivery vehicles [5]. Interestingly, EVs exert many biological processes contributing to health and disease homeostasis [6,7] and cardiovascular disease has become a fast expanding field of interest within EVs.
Atherosclerotic cardiovascular disease (ASCVD) remains the leading cause of death and disability worldwide and atherothrombosis underlies the majority of cardiovascular events [8]. Atherosclerosis chronically develops from early fatty streak to atheromatous plaque formation that when ruptured leads to platelet activation and, ultimately, thrombus formation [9]. We now acknowledge that atherosclerosis is indeed an inflammatory condition [10]. Understanding the clinical benefits of lipid-lowering therapies, such as statins, and other novel pharmacological approaches has provided evidence on the role of inflammation in all the stages of atherosclerosis. Similarly, EVs have also been related to all phases of the atherosclerotic process from initiation to unforeseen thrombotic complications [11]. In the present review, we explore the latest findings on EV-mediated mechanisms and their potential as diagnostic and prognostic biomarkers as well as novel therapeutic targets in inflammatory atherothrombosis.

2. Extracellular Vesicles: Current State and Emerging Concepts

Extracellular vesicles have been recognized as potential bioactive effectors with important roles in cell–cell communication involved in the regulation of a range of biological processes [6,7]. Distinct stimuli can trigger its release by cell activation and apoptosis or senescence under inflammatory, hypoxic, ischemic, or oxidative conditions, and also by mechanical stress such as high blood shear rates [4]. Once secreted, extracellular and pericellular matrix influence EV spreading locally and remotely via the bloodstream [12]. EVs are found ubiquitously in all human extracellular body fluids, such as plasma and urine, and tissues as well as in cell culture conditioned media. Their levels are the result of a fine balance between their shedding and clearance, in which different mechanisms intervene. EVs conform to a highly complex population demonstrated by high heterogeneity in size, biogenesis, origin, content, mechanism of action, and biological function. Indeed, their biogenesis, uptake by target cells and cargo delivery processes (endocytosis, trafficking to endo/lysosomes, membrane fusion, cargo release, and phenotypic changes in the recipient cell) are still unclear (please refer to the comprehensive review for detailed EV basic biology [13]). Mass spectrometry-based proteomics have shown that cargo selection mechanisms are highly regulated by post-translational modifications [14].
The fact that EVs are formed by several but not yet completely known mechanisms convolutes their classification. Following the foundation of the International Society for Extracellular Vesicles (ISEV) one decade ago, the generic ‘extracellular vesicle’ term was adopted to encompass all the distinct names used depending on their source, scale, density, formation, and content [15]. Thus, EVs can be characterized into different subtypes either exosomes, microvesicles (MVs), and apoptotic bodies or small, medium, and large extracellular vesicles, ranging from 20 nm to 5 µm. Briefly, apoptotic bodies (ApoBDs), also known as apoptotic extracellular vesicles (ApoEVs), are large vesicles (>1 µm) with apoptotic nuclear material and permeable membrane formed after cell shrinkage as a result of apoptotic cell death. Despite limited studies on their biological roles, ApoBDs presumably control the transfer of clearance signals among damaged cells. Microvesicles are small particles (100 nm–1 µm) that outwardly bud off from the cellular membrane, predominantly upon cytoskeleton remodeling and externalization of phosphatidylserine (PS), without a nucleus and harboring parental cell-specific surface antigens. They have been also referred to as microparticles, exovesicles, ectosomes, migrasomes, oncosomes, and prostasomes, among others, on the basis of their size and site or context of release. In contrast, nano-sized exosomes (30–100 nm; EXOs), also named exovesicles or endosomes, are endosomal intraluminal vesicles secreted when intracellular multivesicular bodies fuse with the plasma membrane. Multiple pathways have been identified in the regulation of exosome formation [16]. Finally, newly discovered EV subtypes named ‘exomeres’ and ‘supermeres’ have recently revolutionized the field [17,18]. These are described as non-membranous vesicles with a size of ≤50 nm bearing lipid and protein contents. However, there is still debate on whether they should be considered as EVs or extracellular particles like lipoproteins. Given the increasing relevance of EVs, it becomes crucial to uncover the precise function of each vesicle subtype. The following review will specifically report on MVs or large EVs due to their relevant role in ASCVD. Despite the review mainly focusing on MVs, we designate them as EVs throughout the manuscript and only use ‘MV’ or ‘large EVs’ terms to refer to those studies clearly reporting MV-associated data (with evidence of specific markers or subcellular origin). It is important to bear in mind that articles often refer to a mixed population of EVs due to their overlapping size and composition. ApoAB and EXO effects on atherosclerosis are reviewed elsewhere [19,20,21] and will not be the main focus of this review.
One of the most fascinating roles of extracellular vesicles is to engage in cellular crosstalk via the transfer of selective biomolecular cargo to recipient cells in an autocrine, paracrine, or endocrine manner to regulate cell function [22]. The effects of EVs on target cells are highly influenced by the parental cell, stimuli, local environment and, especially, composition. EVs are natural carriers of a highly diverse spectrum of bioactive molecular effectors mostly comprising lipids, proteins, and nucleic acids. In fact, there are several web-based resources and databases available to dig into these vesicular molecules such as Vesiclepedia, ExoCarta, EVpedia, exoRBase, EV-TRACK platform, and exRNA atlas. By means of lipidomics studies, EVs have shown to be highly enriched in cholesterol, eicosanoids, fatty acids, sphingomyelin, glycosphingolipid, phosphatidylserine, and other bioactive lipids, captured from the cytosol during EV formation or synthesized by EV-metabolizing enzymes. Membrane lipids are organized in a bilayer structure with lipid carriers such as fatty acid binding proteins and have the ability to confer a procoagulant surface. Proteomic and enrichment studies have enabled the unveiling of both membrane-bound and intraluminal soluble proteins in EVs. Although a truly selective and universal marker of EV types is not well-defined yet, specific proteins have been used to distinguish EV subtypes such as Annexin A1, identified as specific marker for MVs shed directly from the plasma membrane [23]. Similarly, cell surface-specific markers (cluster of differentiation (CD) markers) are widely used to characterize different EV subpopulations [24]. It has been debated whether these markers are always reflective of their true cell origin or have just attached to the EV membrane. Indeed, recent findings described that EVs themselves transport extracellular proteins interacting with their plasma membrane named protein corona and that EV-associated proteins could also have a pathophysiological relevance and, therefore, should not be considered as ‘contaminants’ of EV preparations [25,26]. Concerning nucleic acids, distinct RNA classes have been found in EVs including functional messenger RNA (mRNA), non-coding RNA (microRNAs (miRNAs), long non-coding RNAs, and circular RNAs) and other regulatory RNA molecules (transfer RNA, small nucleolar RNA, mitochondrial RNA, vault RNA, among others). Valadi et al. were the first to report the existence of mRNAs and miRNAs in EVs and their functional transfer to EV-targeted cells [27]. It has been proposed that cells may use RNA-binding proteins with distinct binding preferences to secrete miRNAs in exosomes [28] or, even, discard unnecessary or inhibitory RNAs. Not only the type of nucleic acid species but also the quantity is crucial for functional impact. Despite large amounts of data over the last years supporting the role of nucleic-acid-associated EVs on cell communication, the specific sorting, transport, and effect of these RNA molecules on EVs is still a poorly understood process that deserves further attention. In this regard, non-vesicular genetic material [29] and EVs could be co-isolated depending on the vesicle isolation procedures, not reflecting true EV-released contents. Moreover, isolation methods influence the yield of recovery of EV-derived RNA [30]. Thus, discrimination of such confounders and reassessment and integration of data is needed prior to assigning any function to EV-associated nucleic acids.
Another layer of complexity in this constantly evolving field includes technical challenges inherent of EVs. The biological, biochemical, and biophysical properties of EVs often overlap between distinct EV subtypes and might be cell-type- and cell-region-specific hampering their separation, discrimination, and analysis [31]. Moreover, preanalytical basic considerations (sample source, phlebotomy procedure, anticoagulant, temperature, storage, etc.) are of utmost importance to avoid contamination of cell-derived vesicles with other non-EV structures and protein aggregates, among others. Equally vital is the correct reporting of methods to ensure reproducibility and standardization. Thus, isolation methods of choice for EV studies (used in combination or even modified) should depend on yield, purity, integrity, concentration, downstream applications, and scientific question, following the recommendations by the Minimal information for studies of extracellular vesicles guidelines ISEV position paper [15]. A methodological consensus framework for the identification and characterization of EVs in cardiovascular medicine studies has very recently been established [32]. While all these obstacles have to be overcome yet, tremendous advances in EV research simultaneously to the emergence of new technologies will undoubtedly reshape our approach to studying extracellular vesicles.

3. EVs as Pathophysiological Drivers of Atherothrombosis

EVs orchestrate the entire spectrum of atherosclerotic disease phases by means of their intercellular and extracellular communication and adhesion between blood and vessel wall [33]. Human atherosclerotic plaques, from initial lesions to advanced stages, contain EVs expressing markers from vascular and blood cells (leukocytes, red blood cells (RBCs), smooth muscle cells, and endothelial cells (ECs)) [34,35,36]. Therefore, it seems plausible that EVs from most cell types of the cardiovascular system participate in the atherosclerotic process. It should be taken into consideration that isolation of EVs from tissues (by homogenization, dissociation, or disruption) preserving their intact biochemical properties is still a challenging procedure that should be further developed. Nevertheless, the landmark study from Leroyer et al. [35] reported that plaque EVs derive mainly from white blood cells (WBCs), demonstrating the existence of an inflammatory milieu. Of interest, plaque EVs favored the balance towards a proinflammatory environment in the culprit lesion [37] and induced T-cell proliferation in a dose-dependent manner, contributing to vascular inflammation and plaque development [36]. A growing body of evidence supports the notion that immuno-inflammation plays a role in the pathogenesis of atherothrombotic disease either by promoting the disease or lesion resolution and vascular repair. Here, we will unravel the immuno-modulatory effect of EVs in the different stages of atherothrombotic disease.

3.1. Atherogenesis

Endothelial dysfunction (ED) characterized by a proatherogenic environment is an early sign of atherosclerosis development [38]. Under pathological conditions, such as risk factors or injury, circulating EVs contribute towards a dysfunctional endothelium [39,40,41,42,43,44,45,46,47,48,49,50]. Beyond the well-studied traditional cardiovascular risk factors, atherothrombosis is spearheaded by multiple risk factors such as clonal hematopoiesis and air pollution [51]. It has recently become apparent that particulate matter (PM) could induce an inflammatory response and EC damage favoring the initial formation of atherosclerotic lesion. EVs derived from PM-exposed alveolar epithelial cells fueled inflammatory signaling via inhibitor of the nuclear factor kappa alpha (IκBα)–nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB)–vascular cell adhesion protein 1 (VCAM1) axis [52]. Nevertheless, EVs shed after PM exposure could also mediate crosstalk between EC-immune cells to cope with the inflammatory assault [53], a response impaired in overweight subjects at increased cardiovascular risk [54].
EVs of several cellular origins (mainly derived from ECs) upon distinct inflammatory stimuli (oxidized low-density lipoprotein (LDL), angiotensin II, hypertension, visceral adipose tissue, and infection) or isolated from diseased patients (circulating EVs) directly cause harmful effects on physiological EC function and vasorelaxation by impairing nitric oxide (NO) bioavailability via regulation of NO, nitric oxide synthase, nicotinamide adenine dinucleotide phosphate (NAPDH) oxidase, and prostacyclin as well as reactive oxygen species (ROS) and endoplasmic reticulum stress [55,56,57,58,59,60,61,62]. RBCs play key roles in vascular homeostasis. Not only altered RBC function but also RBC-derived EVs could reduce NO bioactivity and induce ED [63,64]. Likewise, excessive erythrocytosis exhibited by Andean highlanders also induce ED by means of dysfunctional circulating EVs [65]. Extracellular signal-regulated kinase (ERK) 1 embedded in platelet EVs from a mouse model of diabetes induce ED via intracellular adhesion molecule 1 (ICAM-1) [66]. Senescent EC-derived MVs (eMVs) from ACS patients induce premature ED and thrombogenicity through the angiotensin II–induced NADPH oxidase system [67]. On top of this detrimental vascular effect, EVs are also responsible for endothelial permeability [68,69]. Several works suggest that the disturbance of EC barrier function could also be mediated through apoptosis by EVs [70,71,72,73], whereas other studies ascribe to them a protective role as a way of reducing intracellular levels of caspase-3 [74]. Along the same line, Jansen et al. showed that eMV uptake in ECs via annexin I/PS receptor-p38 signaling protects against apoptosis [75]. Li et al. have demonstrated that endothelial progenitor cell (EPC)-derived EVs inhibit ferroptosis and alleviate atherosclerosis vascular injury via the miR-199a-3p/specificity protein 1 axis [76]. Of interest, EVs in basal conditions or from healthy subjects do not show deleterious effects to the vascular wall [77,78], highlighting that the impact of EVs in vivo is cell source- and context-dependent [79,80,81,82,83].

3.2. Atheroinflammation

An additional key feature of early atherosclerosis is the activation of the endothelium and, subsequently, the adhesion and recruitment of immune cells to the vasculature [84]. Importantly, EVs contribute in a multifaceted fashion to instigate vascular inflammation by increasing levels of adhesion molecules, ROS, and proinflammatory cytokines (e.g., interleukin (IL)-6 and -8) [85,86,87,88]. Specifically, platelet-derived EVs (pEVs) have shown in vitro to enhance the expression of adhesion molecules including E-selectin, ICAM-1, and VCAM-1 [89,90]; thus, favoring the adhesion of monocytes to the inflamed endothelium. Of note, P-selectin-enriched pEVs were found to promote leukocyte–leukocyte interactions leading to monocyte diapedesis through the endothelium even when flow conditions were not favorable [91]. Such chemo-attraction can be regulated by multiple mechanisms such as the transfer of Regulated on Activation, Normal T cell Expressed and Secreted (RANTES) protein to ECs [92] and might contribute to long-term leukocyte differentiation [93]. In this regard, apoptotic-platelet-derived EVs can mediate the polarization towards M2 monocyte and resident macrophages. pEVs not only facilitate leukocyte adherence to the arterial wall but also to the atherosclerotic plaque [94]. In contrast, pEVs can also reshape the secretome of angiogenic early outgrowth cells (EOC) magnifying EOC-driven endothelial regeneration and promoting vascular protection [95]. Latter findings emphasize the importance of the environment in the final EV-directed effect.
A similar effect of chemo-attraction and transmigration of leukocytes to ECs towards the plaque is induced by EVs from other cellular sources [96]. Of note, neutrophil-derived EVs (nEVs) cause myeloperoxidase-mediated damage of vascular ECs [97]. nEVs shuttle Cox-1 substrate arachidonic acid (AA) from neutrophils to platelets, thereby fostering thromboxane A2, EC activation, and neutrophil recruitment in an experimental model of pulmonary hypertension [98]. Thus, nEVs are involved in platelet–neutrophil crosstalk. Leukocyte-derived EVs released upon proinflammatory triggers can adhere to ECs and promote further leukocyte adhesion in a positive feedback loop [99,100,101]. One of the mechanisms by which EVs exert their proinflammatory activity is by harboring mitochondria or mitochondria-derived proteins, the so-called mitovesicles [102,103]. Extracellular mitochondria cause inflammatory responses to injury because they bear damage-associated molecular patterns [104,105], mitochondrial DNA [106], and oxidized proteins [107]. For instance, monocyte-derived mitovesicles can induce type I interferon and tumor necrosis factor (TNF)α response in ECs [108]. It has also been described that coronary artery disease (CAD) patients with low levels of mitochondrial oxidase subunit I in monocyte-derived EVs are at higher risk for new cardiovascular events [109]. pEVs released from thrombin-stimulated platelets also have embedded mitochondria that act as a substrate for phospholipase A2 IIA [110]. Altogether, EVs selectively enriched with mitochondrial material promote the expression of inflammatory mediators and drive leukocyte recruitment in the vasculature. Another mechanism involved in the vascular pro-inflammatory milieu is the exporting and transferring of miRNAs to target cells. Endothelial-derived EVs (eEVs) have shown to transport miRNA-155 and reprogram monocytes into an M1 proinflammatory phenotype [111] and miRNA-92a to macrophages promoting an atheroprone phenotype [112]. Not only miRNAs but also post-translational modifications can influence EV-driven inflammatory signals to ECs. Recently, Yang et al. have shown that monocyte–EC interactions responsible for EC inflammatory phenotype depend on EV-associated glucose transporter 1 and glycosylation [113].
Conversely, EVs can also act as inflammatory repressors. Endothelial-derived EVs bearing the endothelial protein C receptor-activated protein C complex induce protease-activated receptor (PAR) 1-dependent survivable, cytoprotective, and anti-inflammatory effects on ECs [114]. MVs derived from lipid-loaded macrophages (CD16+) and containing LRP5 induce anti-inflammatory phenotypes in macrophages in an autocrine and paracrine fashion [115]. The EC-derived miRNA-143/-145 cluster is delivered to adjacent smooth muscle cells, through a MVs-mediated mechanism, to induce a vasculoprotective phenotype [116]. Neutrophil-derived EVs induce the release of transforming growth factor (TGF) β1 from macrophages with anti-inflammatory effects [117] that could be reinforced by the expression of annexin I on the surface of nEVs [118]. Moreover, ICAM-1 is downregulated by miRNA-222 embedded in eEVs [119]. Another example of a favorable vesicle-mediated miRNA transfer is the anti-inflammatory miRNA-10 targeting monocyte NF-κB pathway [120]. When EVs are taken up by B cells and monocytes, they are able to reprogram target cells towards an anti-inflammatory profile dependent on distinct factors [121]. Continuing with this dichotomy of pro- and anti-inflammatory properties of EVs and in an attempt to untangle the particular contribution of each vesicle subtype to the early phase of atherosclerotic disease, Hosseinkhani et al. provided evidence of a differential immunomodulatory content (inflammatory signals, chemokines, and cytokines) between small and large EVs released from activated ECs [122,123]. Specifically, sEV stimulate monocyte migration while lEVs are prompt to induce ICAM-1 expression and inflammation profile in target cells.
Even though most of the data regarding EVs as pro- and anti-inflammatory mediators in atherosclerosis derives from in vitro models, several studies with EVs from pathological conditions have been conducted [124,125,126,127,128]. Deepening to the mechanism, monocyte-derived EVs carrying IL-1β are involved in EC activation [101] and, consistently, IL-1β-rich eEVs in inflammation in the setting of acute coronary syndrome (ACS) [129]. Remarkably, MVs within atherosclerotic plaques are able to recruit pro-inflammatory cells through transfer of ICAM-1 to ECs [130].
In addition to chronic vascular inflammation, atherosclerosis is also characterized by an impaired resolution response leading to non-resolving inflammation persistence. EVs circulating in plasma could also exert such immunomodulatory potential in atherosclerosis. For instance, deregulated anti-inflammatory and pro-resolving molecules (CD5L and Resolvin E1 bioactive lipid mediator) in circulating EVs contribute to the systemic hyperinflammation in the context of chronic liver disease [131]. In atherosclerotic inflammation, miR-155 was found decreased in a model of atherosclerosis regression and increased in urinary EVs (uEVs) from unstable CAD patients. Interestingly, CAD progression was associated to increased CD45+ and CD11b+ uEVs, and decreased in CD16+ uEVs, pointing at miR-155 and uEVs as a potential prognostic marker of disease progression and severity and a therapeutic target [132].

3.3. Lesion Progression

3.3.1. Atherogenic Foam Cell Formation

The adaptive immune system appears at the center of human ASCVD development. Oxidized (ox-LDL) and aggregated LDL particles are taken up by the macrophages that, by removing the excess of lipids, become foam cells. As a consequence, macrophage lipid accumulation via shedding of EVs (1) inhibits their clearance [133], (2) leads to the formation of oxidation-specific epitopes in a subset of cholesterol-induced EVs [134], and (3) amplifies the inflammatory response. For instance, monocyte-derived EVs promote T-cell infiltration in atherosclerotic plaques [135]. More specifically, macrophages have shown to promote proinflammatory and proatherogenic phenotypes in recipient cells through secretion of EVs conveying miRNAs. Upon lipid uptake, atherogenic (cholesterol-loaded) macrophages secrete EVs and deliver miRNA-146a to naïve macrophages where it represses the expression of specific promigratory target genes (insulin-like growth factor 2 mRNA-binding protein 1 and human antigen R), leading to decreased macrophage migration and potential entrapment in the atherosclerotic plaque [133]. Finally, MV-mediated miR-223 transfer reinforces the macrophage activation loop [136]. Ox-LDL also induces EV release from ECs containing miR-155 which promotes M1 macrophage polarization [111]. Indeed, beyond macrophages and foam cells, ECs and platelets also secrete EVs that regulate macrophage polarization and contribute to atherosclerotic plaque progression. TNFα or low shear stress-induced EC-derived EVs also target macrophages [137]. Activated pEVs stimulate ox-LDL phagocytosis and inflammatory cytokine release by macrophages [138]. Thereafter, macrophages and foam cells undergo apoptosis, either by their phagocytosis of ceramide and AA-rich EVs [70,139] or by the EV-driven transfer of programmed-cell-death-related caspases [140] even in SMCs [73]. Lymphocyte accumulation in inflamed arteries can also display a regulatory and atheroprotective function. B-cell-derived natural IgM recognizes EVs bearing oxidation-specific epitopes and represses their inflammatory signaling and pathological endeavors [141]. Moreover, natural IgM antibodies inhibit MV-driven coagulation and thrombosis [142].

3.3.2. Vascular Smooth Muscle Cell Proliferation, Migration, and Phenotype Switching

During lesion growth, vascular smooth muscle cells (VSMC) located in the media change from a contractile to a proliferative phenotype and migrate into the intima layer. VSMC cell proliferation and migration, another key step in the development of atherosclerosis, is also influenced by EVs either from platelets [143] or foam cells [144]. Platelet-released EVs induce a pro-inflammatory phenotype in VSMC affecting vascular remodeling [143]. In addition, platelet-derived MVs have shown to promote VSMC proliferation and migration via a platelet-derived growth-factor-independent mechanism [145,146] and calcium oscillations and transient receptor potential vanilloid 4 [147], respectively. Macrophages release EVs rich in cholesterol that can be taken up by VSMC [148]. The contribution of macrophage foam cell-derived EVs to enhance VSMC adhesion and migration was shown by the transfer of integrins and subsequent downstream activation of ERK and Ak strain transforming (AKT) [144] and the transport of miR-19b-3p and targeting juxtaposed with another zinc finger gene 1 [149]. Those EVs rich in tissue factor (TF) have the capacity to modulate VSMC migration though PAR interaction [150]. Another study found that VSMC-derived EVs intervene in autocrine VSMC proliferation by upregulation of mitogen-associated protein kinase [151]. Interestingly, other studies address this question with a translational perspective. EVs bearing Ras-related protein 1 from patients with metabolic syndrome favored VSMC migration and proliferation [152]. Plasma EV containing miRNA-501-5p promotes VSMC phenotypic modulation through targeting Suppressor of mothers against decapentaplegic homolog (Smad) 3 in patients with in-stent restenosis one year after coronary stent implantation [153]. Conversely, several studies emphasize the fact that EVs and their miRNA content present an atheroprotective effect on VSMCs. In this regard, miRNA-223-EVs inhibited VSMC migration and proliferation, thereby decreasing plaque size [154]. Similarly, both high shear stress and Krüppel-like factor 2-expressing ECs-derived EVs enriched in the miR-143/145 cluster prevented VSMC dedifferentiation [116]. Moreover, macrophage-derived EVs also transmit miR-503-5p into VSMCs and inhibit their proliferation and migration via the Smad7-Transforming growth factor-β axis [155]. Finally, MVs reduced VSMC proliferation, migration and subsequent neointima formation by delivering functional miR-126 into recipient VSMCs [155]. Even EVs from adipose mesenchymal stem cells were involved in the regulation of VSMC phenotype to limit intimal hyperplasia [156].

3.3.3. Intravascular Calcification

Under pathological stimuli, EVs released by VSMCs and macrophages aggregate between collagen fibers and serve as a platform for ectopic mineralization eliciting micro- and macrocalcifications in the vessel wall. Vascular calcification in atherosclerosis shows a bimodal behavior in terms of risk and impact; while it has been shown that microcalcifications have a burdening effect on plaque rupture due to the low number of inflammatory cells and being more prone to rupture, higher amounts of calcification within stable plaques at various phases of atherosclerosis progression have also been found. Cellular-derived EVs from VSMCs, ECs, and platelets indisputably regulate the loss of the VSMC contractile phenotype and impact vascular calcification [157,158,159] favoring plaque calcification and atherogenesis [160,161,162,163]. EVs from dysfunctional endothelium [164] and senescent ECs [158,165] promote vascular calcification. VSMCs also release calcifying EVs themselves [166,167]. One described mechanism driving the VSMCs towards calcification involves the regulation of calcifying EVs by sortilin [166]. Oxidative stress also intervenes in this process. VSMC phenotype switching caused an increased Nox5 expression that was responsible for the subsequent extracellular calcium entry, ROS production, reduced contractile phenotype, and enhanced EV release fostering VSMC calcification [168].

3.3.4. Necrotic Core

EV-mediated cell death [73,139] triggers necrotic core formation and, in turn, lipid-rich necrotic core contains high amounts of thrombogenic EVs as well [35,36,169], conferring a procoagulant potential and instability to the atherosclerotic lesion and shifting towards a vulnerable plaque.

3.4. Advanced Lesion and Plaque Rupture

Atherosclerotic plaque erosion or rupture and the subsequent thrombosis leads to, eventually, acute ischemic syndromes. The interdependence between innate immune cells and platelets is essential for plaque destabilization and thrombotic occlusion [170,171].

3.4.1. Arterial Neoangiogenesis and Intraplaque Hemorrhage

During the last steps of atherosclerosis development, an increased number of vasa vasorum brings about intraplaque hemorrhage, a highly inflammatory environment, and a rapid activation of the coagulation cascade and platelet aggregation forming a fibrin mesh [172,173]. Several studies support the notion that EVs exert regulatory roles in these events [174,175]. EPC-EVs activate ECs and stimulate angiogenesis [176]. Inflammation-induced EC-derived EVs enhanced vascular endothelial growth factor B expression in pericytes, which in turn might mediate pathological neovascularization and vascular leakage in response to an inflamed endothelium [177]. Atherosclerotic EC-released EVs promoted angiogenic responses in ECs by a thrombospondin-1-depedent mechanism [112]. Communication between ECs and immune cells plays key roles in disease development. Spleen EC-derived VCAM1+- EVs mobilize splenic monocytes in myocardial infarction (MI) [178]. Our group has reported that hypoxic ECs release endothelial-cell-derived MVs rich in TF and transport miRNA-126 that induces monocyte recruitment into the ischemic zone, reprogramming of monocytes and their differentiation into EC-like cells, and angiogenic response, which altogether improve tissue reperfusion [179,180]. Of note, TF-containing eMVs could interact via paracrine signaling with other microvascular ECs and trigger angiogenesis ex vivo and postischemic collateral vessel growth in vivo [181]. MVs from human atherosclerotic plaques promoted in vivo angiogenesis [182]. Of interest, in vitro studies showed that the mechanism was mediated by MVs expressing the CD40 ligand [182]. Similarly, MVs from ischemic muscle induced postnatal vasculogenesis [183]. Fibrinolytic activity of EC and WBC-derived EVs could also favor angiogenesis and hamper thrombus dissolution [184]. Furthermore, Loyer et al. found that small and large cardiomyocyte-derived EVs are locally released from the heart to regulate the inflammatory process in a mouse model of myocardial ischemia [185]. Indeed, ischemia-induced cardiosomes stimulate cardiac angiogenesis both in vitro and in vivo [186]. Besides intraplaque neovascularization and bleeding, plaque, EC, and immune-cell-derived EVs influence endothelium denudation and fibrous cap integrity. EVs exert structural extracellular matrix degradation and fibrous cap weakening via metalloprotease activities [37,187,188,189,190,191,192,193]).

3.4.2. Thrombus Formation after Rupture

Fibrous cap disruption within the atherosclerotic plaque results in the exposure of subendothelial extracellular matrix (collagen and von Willebrand factor) and thrombogenic substances and, thus, a huge release of tissue factor triggering platelet activation, aggregation, and thrombus formation in the damaged area [194]. EVs from platelet/monocyte aggregates are capable of modulating human atherosclerotic plaque reactivity [195]. Under in vitro simulated conditions of coagulopathy, EV from different cellular origins (EC and platelet) and at distinct concentrations showed a divergent but procoagulant effect within the coagulation process [196]. Numerous studies have dissected in vitro and in vivo the procoagulant effect of EVs, mainly MVs, leading to thrombosis [142,197,198,199,200,201,202,203,204,205,206,207]. Procoagulant EVs within the necrotic core of atherosclerotic plaques provoke the burst of the coagulation cascade upon plaque rupture [169] being TF+-MVs the main initiators [208,209,210]. TF is functionally transferred to monocytes through MVs [211] in a PAR2-filamin A-mediated process [212]. The fact that the proinflammatory cytokine IL-33 [213] and EC-derived EVs [214] trigger the release of procoagulant TF+-MVs underscores the synergy between inflammation and thrombosis. TF+-MVs are also engulfed by platelets and trigger platelet aggregation [215,216]. In addition, TF-rich monocyte-derived EVs are recruited at the thrombus site [217]. TF shares the limelight of plaque thrombogenicity with phosphatidylserine. Negatively charged PS exposure in EV surface confers procoagulant activity [218], enables the concentration of factors VII/VIIa in the EV membrane [219], enhances platelet adhesion to the endothelium, and it is important for TF function, thrombin generation, and clot formation [220]. Beyond these two key factors (TF and PS), EVs were shown to directly modulate the clotting. A seminal study on leukocyte-derived EVs expressing P-selectin glycoprotein ligand-1 (PSGL-1) on their surface deconvolutes the crosstalk between platelets and TF-rich immune cells at the lesion site. Interactions between platelet adhesion molecule P-selectin and PSGL-1+-EVs allowed higher recruitment and activity of TF perpetuating thrombus formation [221]. Not only WBC-derived EVs participate in this dynamic process, circulating and platelet-derived MVs also enhance thrombus formation and growth on thromboactive substrates [205]. We also found that pEVs bearing surface epitopes of platelet adhesion and activation in perfused blood tend to bind more to adhered platelets and prothrombotic surfaces stimulating further platelet deposition and thrombus growth [207]. Interaction of platelets with subendothelial matrix, immobilized platelet surfaces [222], or fibrin [223] depends on pEV-associated αIIbβ3a integrin [224].
Several other platelet biological components contribute to this phenomenon such as protein disulphide isomerase [225], factors VIII and Va [226,227], and bioactive lipids [228]. Platelet-derived EVs also modulate EC–monocyte interactions [89] likely inducing enhanced expression of cell adhesion molecules under high shear stress conditions [90]. Interestingly, pEVs support megakaryocyte differentiation and platelet production that is relevant to keep a physiological level of circulating platelets after their consumption during thrombosis [229]. Recently, we have investigated the proteostatic characteristics of EVs shed from platelets upon thrombin activation by an untargeted proteomic approach [230]. We have identified proteins involved in cell–cell interaction and signaling events underlying thrombus formation such as CUB domain-containing protein-1 or membrane glycoprotein gp140, fermitin family homolog 3 or kindlin-3, and the novel pEV-associated protein protocadherin-α4, among others. Deciphering platelet and cell signaling and biology by proteomics technique [230,231] will help to unveil paracrine regulation of EVs and provide new therapeutic targets involved in occlusive thrombus formation and atherothrombosis.

4. Extracellular Vesicles as Diagnostic and Prognostic Biomarkers

EVs are released in healthy physiological steady conditions. Clinical ASCVD development is associated with high release of EVs. Assessment of vascular status is crucial for atherosclerosis prevention and to this aim EVs can serve as potential biomarkers in the clinics [38,232]. Molecular footprints of EVs provide clues about their cellular origin under healthy and diseased states, configuring a precious biomedical tool for liquid biopsy, easily detectable in body fluids. Changes in circulating EVs of different cellular origin have been widely reported for almost all cardiovascular risk factors [233,234,235,236] and pathologies [237,238,239] reflecting their pathophysiological severity. Beyond diagnostic use, EVs have been suggested for the prediction of major adverse cardiovascular events (MACE) [230,231,232,233,234,235,236,237,238,239,240,241,242,243] and mortality [244]. Circulating procoagulant MVs were found at higher levels in patients with ACS than healthy subjects [238,239,240,245,246,247,248,249,250,251,252,253,254,255]. Table 1 lists representative studies showing EV changes in cohorts of patients with distinct ASCVD risk factors and disorders with potential utility as diagnostic and prognostic biomarkers.
Among different cell-derived EVs, EVs from endothelium stand out due to their capacity to mirror ED [256,257], correlate to cardiovascular outcomes [258,259], and predict disease severity [260]. Circulating EVs associate to ED and plaque rupture in ACS patients [261]. Associations between circulating EVs and angiographic lesions have also been reported [262,263]. Circulating EV levels correlated with a higher calculated 10-year Framingham CAD risk [264]. Risk assessment is critical in primary prevention of ASCVD, especially for asymptomatic patients with subclinical atherosclerosis. In this context, EVs emerge as promising biomarkers for CV risk stratification. We found that patients with familial hypercholesterolemia (FH) presenting subclinical lipid-rich atherosclerotic plaques have significantly higher amounts of circulating lymphocyte-derived (CD3+/CD45+) EVs than FH patients with fibrous plaques [50] in agreement with other studies that reported high levels of EVs from T lymphocytes in subjects with essential hypertension [235] and about to develop a MACE [265]. Thus, lymphocyte-derived EVs might help to map atherosclerotic plaque burden. Accordingly, increased circulating levels of leukocyte-derived EVs in patients with unstable atherosclerotic plaques signal for plaque vulnerability [246]. Furthermore, we also found that asymptomatic high cardiovascular risk FH patients have significantly higher number of circulating prothrombotic platelet EVs [266]. Patients with FH, despite being treated according to guidelines, have ongoing innate immune cell and platelet activation. Indeed, in asymptomatic high cardiovascular risk FH patients from the same cohort, pan-leukocyte-derived and activated neutrophil-derived circulating EVs as well as EVs bearing markers of platelet activation were significantly increased in patients about to suffer an ischemic event [242]. This specific EV signature is a prognostic marker able to improve the prediction of clinical events if included in the risk factor models. Altogether, it is clear that multiple biomarker-based strategies for patient stratification including EVs hold promise in precision medicine.
Table 1. Extracellular vesicles (EVs) as potential markers of ASCVD diagnosis and prognosis. Studies showing EV changes in distinct ASCVD risk factors and pathologies.
Table 1. Extracellular vesicles (EVs) as potential markers of ASCVD diagnosis and prognosis. Studies showing EV changes in distinct ASCVD risk factors and pathologies.
Clinical Condition/ContextEV SubtypeCargoReference
Cardiovascular risk factors↑ eEV (CD144+, CD105+) (Hypertension) [267]
↑ lEV (CD3+) (Hypertension) [235]
↑ eEV (CD31+/42; CD31+/AnnV+), pEV (CD31+/42+) (Diabetes)miR-126[236]
↑ lEV (CD45+), nEV (CD15+), pEV (CD62P+) (Familial hypercholesterolemia [FH]) [242]
↑ pEV (CD42a+, CD62P+) (Obesity) [268]
↑ pEV (CD41+), eEV (CD62E+), lEV (CD45+) (Smoking)CD40L[269]
Pulmonary hypertension↑ eEV (CD31+, CD144+, CD62E+), lEV (CD45+) [260]
Endothelial dysfunctioneEV (CD31+/CD42b) [255]
eEV (CD31+/AnnV+) [256]
FH and subclinical atherosclerosis↑ lEV (CD3+/CD45+) [50]
↑ pEV (TSP1+/CD142+)Tissue factor[266]
Coronary artery disease (CAD)↑ CD31+/Annexin V+ EV [258]
↑ eEV (CD31+), pEV (CD42b+)miR-126, miR-199a[241]
↑ eEV (CD31+/AnnV+) [256]
↑ eEV (CD31+ and CD51+) [238]
↑ eEV (CD31+) [270]
Arterial hypertension and coronary artery disease↑ eEV (CD31+/41, CD62E+, CD144+) [271]
Diabetes and CAD↑ eEV (CD62E+, CD31+)↓LAP(TGF-β1), PD-ECGF, PF4, TSP1[233]
Coronary heart disease↑ eEV (CD31+/CD42, CD144+) [272]
↑ pEV (CD41+), eEV (CD144+) [273]
↑ eEV (CD31+ and CD146+) [248]
Stable angina ↑ eEV (CD31+), pEV (CD41+) [274]
Acute coronary syndrome↑ pEV (CD31+, CD41a+) [275]
↑ eEV (CD146+) [276]
↑ eEV (CD144+), ErEV (CD235a+), pEV (CD41a+) [261]
Myocardial infarction (MI)↑ CM-EV and eEV (CD31+/CD41)Caveolin-3, TnT[185]
↑ ErEV (CD235a+) [207,252]
↑ eEV (CD144+) [277]
↑ l/mEV, eEV, ErEV, activated vascular cell-EV, (CD66b+/62E+/142+ for coronary thrombosis) [245]
↑ eEV (CD154+, CD62E+), pEV (CD62P+) [278]
↑ eEV (CD31+), pEV (GPIbα) [247]
↑ eEV (CD31+, CD146+), pEV (CD42b+) [254]
↑ pEV (CD61+), mEV (CD14+)Tissue factor[244]
↑ lEV(CD45+), pEV (CD61/62P+), ErEV (CD235a+)Lactadherin, Fbn[279]
↑ eEV (CD144+), pEV (CD62P+/41), mEV (CD14+)Tissue factor[280]
MI and stable angina↑ pEV (CD51+/61+), eEV (CD42/31+), mEV (CD14+)Tissue factor[250]
MI in chronic kidney disease↑ eEV (CD154+, CD62E+), pEV (CD62P+) [278]
Ischemic cerebrovascular disease↑ eEV (CD144+, CD31+, CD62E+/41a, AnnV+) [281]
↑ eEV (CD62E+, CD31+/, CD42b and Annexin V+) [237]
Stroke↑ NPC-EV (CD34+, CD56+) [282]
↑ pEV (CD41+) [283]
↑ eEV (CD62E+) [259]
↑ eEV (CD31+, CD144+, CD146+), lEV (CD45+) [243]
Peripheral arterial disease↑ pEV (CD41+/61+)Calprotectin[284]
↑ cEVSonic hedgehog[285]
↑ eEV (CD144+) [286]
Venous thromboembolism ↑ eEV (CD62E+), mEV (CD14+)PSGL-1+[287]
Atrial fibrillation and stroke ↑ eEV (CD41/31+), pEV (CD61+) [288]
Autoimmune chronic inflammatory diseases ↑ eEV and pEV [289]
AnnV indicates annexin V; cEV, circulating extracellular vesicle (EV); CD, cluster of differentiation; CM-EV, cardiomyocyte-derived EVs; eEV, endothelial-derived EV; ErEV, erythrocyte-derived EV; Fbn, fibrinogen; LAP(TGF-β1), latency-associated peptide (transforming growth factor β1); lEV, leucocyte derived EV; mEV, monocyte-derived EV; nEV, neutrophil-derived EV; NPC-EV, neural precursor cells; PD-ECGF, platelet-derived endothelial cell growth factor; PF4, platelet factor 4; pEV, platelet-derived EV; PSGL-1, P-selectin glycoprotein ligand-1-positive; TnT, troponin T; and TSP1, thrombospondin 1; ↑, increase.
When studying circulating EVs as markers of disease it is relevant to consider confounders. We know that EV levels can be impacted by sex [290], age [291], pregnancy [292], a high-fat meal [293], and exercise intensity [294,295,296]. Indeed, skeletal muscle myofibers are a major source of EVs, more than white adipose tissue, that can reach the circulation in vivo [297]. Large clinical trials to fully demonstrate EV potential are an indispensable pre-requisite before entering the clinical arena.

Pharmacological Modulation of Extracellular Vesicles

EVs are amenable to pharmacological intervention. Several drugs have shown the ability to influence EV shedding as well as EV composition. For instance, antioxidants [298], antihypertensive drugs (angiotensin II receptor antagonists [299] and calcium channel blockers [300]), lipid-lowering therapy (statins [301] and ezetimibe [302]), and antiplatelet drugs (acetylsalicylic acid [303], GPIIb/IIIa inhibitors [304], clopidogrel [305], and ticlopidine [306]) have shown to reduce EV levels, as reviewed elsewhere [4]. Recently, new antidiabetic drugs have been studied in this regard. Circulating EVs from CAD patients via the angiotensin signaling induce higher expression of sodium-glucose cotransporter (SGLT)-2 protein promoting endothelial senescence and dysfunction as well as pro-atherothrombotic responses in coronary ECs; effects that can be targeted by SGLT-2 inhibitors [307]. In a similar fashion, thoracic adipose tissue secretes ceramide 16:0 metabolite within EVs that enhance vascular redox state in the obesity setting, a phenomenon that is amenable to treatment with glucagon-like peptide 1 receptor antagonist liraglutide [308]. Liraglutide ameliorated pro-atherogenic eMV release in women with gestational diabetes, highlighting atheroprotective effects of liraglutide [309]. Importantly, the anti-inflammatory drug colchicine, which has demonstrated beneficial effects on MACE after MI [310] and in patients with chronic coronary disease [311], inhibits NLR family pyrin domain containing 3 (NLRP3) inflammasome and reduces EV-associated NLRP3 protein levels [312].

5. Extracellular Vesicles as Therapeutic Tools

EVs are considered as transfer systems capable of delivering biological cargoes in health and disease (Figure 1). Given their beneficial effects on vascular homeostasis and atheroprotection and their recognition as the next generation of cell-derived therapeutics [313], EVs show promising perspectives for ASCVD therapy as has recently been suggested for pEVs [314]. EVs offer multiple advantages as drug delivery vehicles such as stability in the bloodstream, biodegradability, ability to cross biological barriers and target specific cells or tissues, protection to their loaded content, low immunogenicity (if the same patient’s sample is used), among others [315]. In cardiovascular medicine applications, we could add the fact that their biodistribution as intravenously-injected EVs could reach the whole vascular network. As major impediments and due to multiple functional properties displayed by EVs (both beneficial and detrimental), tight control of their ultimate mission as well as cell and tissue specificity should be stably achieved to avoid off-target effects [316].
Up to now, one of the areas where the therapeutic potential of EVs has raised interest and has been widely explored is in cardiac regenerative therapies. EVs from distinct cell types have shown to impact on cardiac cells providing a tissue-reparative phenotype. pEVs induced VEGF-dependent revascularization after chronic cardiac ischemia [317]. Concerning EVs from heart immune cells, dendritic-cell-derived EVs infiltrate the ischemic myocardium and act as activators of CD4+-T-cells and improve wound healing post-MI [318]. Macrophages-derived EVs enhance inflammation in injured cardiac cells [319]. Inflammation-induced EC-derived EVs regulated the expression of angiogenic genes in pericytes [320], indicating the crosstalk between cardiac cells and pericytes in myocardial remodeling. EVs have recently gained attention in therapeutics as paracrine contributors to reparative functions of stem cells (mainly from endothelial progenitor or mesenchymal (MSC), cardiac, embryonic, and induced pluripotent stem cells) in the heart. The vesicular fraction of conditioned media from hypoxic MSC was able to reduce infarct size and oxidative stress and improve cardiac function, thereby potentiating myocardial viability and preventing further damage to the heart after MI [321,322]. In addition, stem- and progenitor-derived EVs injected intramyocardially and intravenously in preclinical murine models of myocardial ischemia/reperfusion injury, were able to improve cardiac function by modulating monocyte and macrophage activity and their polarization into anti-inflammatory [323].
miRNA-transferring EVs have also recently been considered as therapeutic vectors following myocardial ischemia [324]. miRNA-126 MVs stimulated re-endothelization following vascular injury by targeting Sprouty-related EVH1 Domain Containing-1 protein [325]. Upon the inflammatory insult, ECs release VCAM-1+–EVs that promote splenic monocyte mobilization and transcription activation in acute MI [178]. Therapeutic targeting of the EV-associated VCAM-1 axis after AMI may specifically limit splenic monocyte cell movement and modulate the inflammatory phase of heart injury, thereby improving cardiac function and long-term prognosis. In fact, therapeutic immunomodulation of the neutrophil response to MI using EV vectors (EC-derived EVs containing VCAM1 and miR-126) has already been tested [326].
Specific to atherosclerosis, T-lymphocytes-derived MVs bearing sonic hedgehog induce neoangiogenesis and endothelial repair through the NO pathway [327,328]. Of interest, a very recent study has shown that EVs loaded with Smad2/3-siRNA from subcutaneous adipose tissue stem cells and bone marrow MSCs ameliorate vascular dysfunction and regress inflammation-mediated atherosclerosis [329]. Finally, EVs from human placental-expanded stromal cells bear a functional corona with immunomodulatory and angiogenic effects with potential application for peripheral artery disease [330].
Engineering of nanoparticles with encapsulated biological molecules as novel therapeutic vehicles is another emerging option. These nanoparticles can be engineered with specific surface molecules to boost their in vivo stability, specificity, and bioactivity [331,332,333,334]. Among the increasing number of studies in this regard (mainly focused on cancer), it is relevant to highlight the magnetic nanoparticle-based gene transfer into the vascular endothelium in order to tackle the diseased vasculature [335,336,337]. To increase delivery efficiency to specific tissues, EVs have been combined with biomaterial scaffolds (tissue engineering). For instance, loading EC-derived EVs with bioprinted 3D structures [338] or a sustained delivery of EPC-derived EVs from injectable hydrogels [339] to support neovascularization and ameliorate cardiac dysfunction after MI. Recently, a synthetic nanotechnology leveraged to mimic the procoagulant function (adhesion and aggregation) of native platelets on nanoparticles has been reported. Platelet-mimicking procoagulant nanoparticles (PPNs), containing PS on their surface, can resist clot lysis and increase clot stability, leading to reduced bleeding and improved survival in rodent models of thrombocytopenia and traumatic hemorrhage. The design of PPNs able to amplify thrombin and fibrin generation at injured sites and enhance homeostatic clot formation emphasizes the multiple therapeutic applications of engineered EVs [340].
There are already clinical studies registered to use EVs in a wide range of diseases (104 trials listed on ClinicalTrials.gov, US National Library of Medicine, Bethesda, MD, USA). Several parameters need to be further explored before the therapeutic effects of EVs observed in the experimental setting can be translated into the clinics such as administration route (systemic injection, topical or radiologic, ultrasound, or endoscopy-guided administration, intranasal injection, oral treatment), dose, immunogenicity, in vivo biodistribution, half-life, and efficiency of cargo delivery [313]. Additionally, standardized, scalable, reproducible, and efficient protocols for EV production with minimal batch-to-batch product variation are required for future engineered EV-based therapeutics [341].

6. Future Perspectives and Conclusions

Extracellular vesicles play essential cell communication roles in cardiovascular biology. Considering the implication of extracellular vesicles in inflammatory atherothrombosis, targeting of the EV-driven innate and immune responses offers new therapeutic avenues. We also pinpointed pathways by which EVs could act as protective factors inducing anti-inflammation and atheroprotection that may serve as potential therapeutic targets against atherosclerosis onset and progression. Further experiments are required in the pathophysiological context, taking into account surrounding cells and spatiotemporal dynamics by using novel imaging techniques [342], omics-based methodology [230], organ-on-chip systems [343,344], or model organisms such as zebrafish [345] to reveal new features of EVs. While challenges remain, EVs emerge as liquid-biopsy-based markers with potential diagnostic and prognostic use in the clinics. Improvements in single-cell and high-resolution flow cytometry will allow better understanding of EVs as biomarkers. Finally, the use of native or bio-engineered EVs might represent novel drug delivery tools to tackle cardiac regeneration, inflammation, and atherosclerosis. We expected that in the near future more preclinical and clinical studies within EVs will help to define the added value of EVs into precision medicine and translate into broader EV-based therapeutic armamentarium for patients with atherosclerotic cardiovascular disease.

Author Contributions

Conceptualization, L.B. and R.S.; writing—original draft preparation, R.S.; writing, M.F.G.; review and editing, R.S., T.P. and L.B.; supervision, L.B.; All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Spanish Ministry of Science and Innovation MCIN/AEI/10.13039/501100011033, Plan Nacional Proyectos Investigación Desarrollo [PID2019-107160RB-I00 to L.B.]; Centro de Investigación Biomedica en Red Cardiovascular [CIBERCV-CB16/11/00411 to L.B.]; Institute of Health Carlos III (ISCIII) [FIS PI19/01687 to T.P.]; cofounded by Fondo Europeo de Desarrollo Regional (FEDER) “Una Manera de Hacer Europa”; Generalitat of Catalunya, Secretaria d’Universitats i Recerca, Departament d’Economia i Coneixement [2017 SGR 1480 to L.B.]. R.S. is the recipient of a Beatriu de Pinós fellowship [2019BP00211] from University and Research Grants Management Agency (Government of Catalonia) co-funded by COFUND- Marie Skłodowska-Curie Actions in the Horizon 2020 programme (European Commission, contract number 801370).

Acknowledgments

We thank the Fundación Investigación Cardiovascular-Fundación Jesus Serra, Barcelona, Spain, for their continuous support.

Conflicts of Interest

R.S. and M.F.G. declare no conflict of interest. T.P. is a founder and shareholder of Glycardial Diagnostics SL and Ivestatin Therapeutics SL (all outside of this work). L.B. received a research grant from AstraZeneca; holds advisory board work for Sanofi, Bayer, and AstraZeneca; received speaker fees from Lilly, MSD-Boehringer, and AstraZeneca; and is a founder and shareholder of Glycardial Diagnostics SL and Ivestatin Therapeutics SL (all outside of this work).

References

  1. Chargaff, E.; West, R. The biological significance of the thromboplastic protein of blood. J. Biol. Chem. 1946, 166, 189–197. [Google Scholar] [CrossRef]
  2. Wolf, P. The nature and significance of platelet products in human plasma. Br. J. Haematol. 1967, 13, 269–288. [Google Scholar] [CrossRef] [PubMed]
  3. Couch, Y.; Buzas, E.I.; Vizio, D.D.; Gho, Y.S.; Harrison, P.; Hill, A.F.; Lotvall, J.; Raposo, G.; Stahl, P.D.; Thery, C.; et al. A brief history of nearly EV-erything—The rise and rise of extracellular vesicles. J. Extracell Vesicles 2021, 10, e12144. [Google Scholar] [CrossRef] [PubMed]
  4. Badimon, L.; Suades, R.; Arderiu, G.; Pena, E.; Chiva-Blanch, G.; Padro, T. Microvesicles in Atherosclerosis and Angiogenesis: From Bench to Bedside and Reverse. Front. Cardiovasc. Med. 2017, 4, 77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. De Freitas, R.C.C.; Hirata, R.D.C.; Hirata, M.H.; Aikawa, E. Circulating Extracellular Vesicles As Biomarkers and Drug Delivery Vehicles in Cardiovascular Diseases. Biomolecules 2021, 11, 388. [Google Scholar] [CrossRef]
  6. Yates, A.G.; Pink, R.C.; Erdbrugger, U.; Siljander, P.R.; Dellar, E.R.; Pantazi, P.; Akbar, N.; Cooke, W.R.; Vatish, M.; Dias-Neto, E.; et al. In sickness and in health: The functional role of extracellular vesicles in physiology and pathology in vivo: Part II: Pathology. J. Extracell Vesicles 2022, 11, e12190. [Google Scholar] [CrossRef]
  7. Yates, A.G.; Pink, R.C.; Erdbrugger, U.; Siljander, P.R.; Dellar, E.R.; Pantazi, P.; Akbar, N.; Cooke, W.R.; Vatish, M.; Dias-Neto, E.; et al. In sickness and in health: The functional role of extracellular vesicles in physiology and pathology in vivo: Part I: Health and Normal Physiology. J. Extracell Vesicles 2022, 11, e12151. [Google Scholar] [CrossRef]
  8. Timmis, A.; Vardas, P.; Townsend, N.; Torbica, A.; Katus, H.; De Smedt, D.; Gale, C.P.; Maggioni, A.P.; Petersen, S.E.; Huculeci, R.; et al. European Society of Cardiology: Cardiovascular disease statistics 2021. Eur. Heart J. 2022, 43, 716–799. [Google Scholar] [CrossRef]
  9. Libby, P.; Buring, J.E.; Badimon, L.; Hansson, G.K.; Deanfield, J.; Bittencourt, M.S.; Tokgozoglu, L.; Lewis, E.F. Atherosclerosis. Nat. Rev. Dis. Primers 2019, 5, 56. [Google Scholar] [CrossRef]
  10. Libby, P. Inflammation during the life cycle of the atherosclerotic plaque. Cardiovasc. Res. 2021, 117, 2525–2536. [Google Scholar] [CrossRef]
  11. 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] [PubMed]
  12. Lewin, S.; Hunt, S.; Lambert, D.W. Extracellular vesicles and the extracellular matrix: A new paradigm or old news? Biochem. Soc. Trans. 2020, 48, 2335–2345. [Google Scholar] [CrossRef] [PubMed]
  13. Van Niel, G.; Carter, D.R.F.; Clayton, A.; Lambert, D.W.; Raposo, G.; Vader, P. Challenges and directions in studying cell-cell communication by extracellular vesicles. Nat. Rev. Mol. Cell Biol. 2022. [Google Scholar] [CrossRef] [PubMed]
  14. Morales-Sanfrutos, J.; Munoz, J. Unraveling the complexity of the extracellular vesicle landscape with advanced proteomics. Expert Rev. Proteom. 2022, 1–13. [Google Scholar] [CrossRef]
  15. Thery, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): A position statement of the International Society for Extracellular Vesicles and update of the MISEV2014 guidelines. J. Extracell Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef] [Green Version]
  16. 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]
  17. Zhang, H.; Freitas, D.; Kim, H.S.; Fabijanic, K.; Li, Z.; Chen, H.; Mark, M.T.; Molina, H.; Martin, A.B.; Bojmar, L.; et al. Identification of distinct nanoparticles and subsets of extracellular vesicles by asymmetric flow field-flow fractionation. Nat. Cell Biol. 2018, 20, 332–343. [Google Scholar] [CrossRef]
  18. Zhang, Q.; Jeppesen, D.K.; Higginbotham, J.N.; Graves-Deal, R.; Trinh, V.Q.; Ramirez, M.A.; Sohn, Y.; Neininger, A.C.; Taneja, N.; McKinley, E.T.; et al. Supermeres are functional extracellular nanoparticles replete with disease biomarkers and therapeutic targets. Nat. Cell Biol. 2021, 23, 1240–1254. [Google Scholar] [CrossRef]
  19. Caruso, S.; Poon, I.K.H. Apoptotic Cell-Derived Extracellular Vesicles: More Than Just Debris. Front. Immunol. 2018, 9, 1486. [Google Scholar] [CrossRef] [Green Version]
  20. Heo, J.; Kang, H. Exosome-Based Treatment for Atherosclerosis. Int. J. Mol. Sci. 2022, 23, 1002. [Google Scholar] [CrossRef]
  21. Paone, S.; Baxter, A.A.; Hulett, M.D.; Poon, I.K.H. Endothelial cell apoptosis and the role of endothelial cell-derived extracellular vesicles in the progression of atherosclerosis. Cell Mol. Life Sci. 2019, 76, 1093–1106. [Google Scholar] [CrossRef] [PubMed]
  22. Joshi, B.S.; de Beer, M.A.; Giepmans, B.N.G.; Zuhorn, I.S. Endocytosis of Extracellular Vesicles and Release of Their Cargo from Endosomes. ACS Nano 2020, 14, 4444–4455. [Google Scholar] [CrossRef] [Green Version]
  23. Jeppesen, D.K.; Fenix, A.M.; Franklin, J.L.; Higginbotham, J.N.; Zhang, Q.; Zimmerman, L.J.; Liebler, D.C.; Ping, J.; Liu, Q.; Evans, R.; et al. Reassessment of Exosome Composition. Cell 2019, 177, 428–445.e18. [Google Scholar] [CrossRef] [Green Version]
  24. Dickhout, A.; Koenen, R.R. Extracellular Vesicles as Biomarkers in Cardiovascular Disease; Chances and Risks. Front. Cardiovasc. Med. 2018, 5, 113. [Google Scholar] [CrossRef] [PubMed]
  25. Palviainen, M.; Saraswat, M.; Varga, Z.; Kitka, D.; Neuvonen, M.; Puhka, M.; Joenvaara, S.; Renkonen, R.; Nieuwland, R.; Takatalo, M.; et al. Extracellular vesicles from human plasma and serum are carriers of extravesicular cargo-Implications for biomarker discovery. PLoS ONE 2020, 15, e0236439. [Google Scholar] [CrossRef] [PubMed]
  26. Toth, E.A.; Turiak, L.; Visnovitz, T.; Cserep, C.; Mazlo, A.; Sodar, B.W.; Forsonits, A.I.; Petovari, G.; Sebestyen, A.; Komlosi, Z.; et al. Formation of a protein corona on the surface of extracellular vesicles in blood plasma. J. Extracell Vesicles 2021, 10, e12140. [Google Scholar] [CrossRef]
  27. Valadi, H.; Ekstrom, K.; Bossios, A.; Sjostrand, M.; Lee, J.J.; Lotvall, 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] [Green Version]
  28. Shurtleff, M.J.; Temoche-Diaz, M.M.; Karfilis, K.V.; Ri, S.; Schekman, R. Y-box protein 1 is required to sort microRNAs into exosomes in cells and in a cell-free reaction. eLife 2016, 5, e19276. [Google Scholar] [CrossRef]
  29. Arroyo, J.D.; Chevillet, J.R.; Kroh, E.M.; Ruf, I.K.; Pritchard, C.C.; Gibson, D.F.; Mitchell, P.S.; Bennett, C.F.; Pogosova-Agadjanyan, E.L.; Stirewalt, D.L.; et al. Argonaute2 complexes carry a population of circulating microRNAs independent of vesicles in human plasma. Proc. Natl. Acad. Sci. USA 2011, 108, 5003–5008. [Google Scholar] [CrossRef] [Green Version]
  30. Gutierrez Garcia, G.; Galicia Garcia, G.; Zalapa Soto, J.; Izquierdo Medina, A.; Rotzinger-Rodriguez, M.; Casas Aguilar, G.A.; Lopez Pacheco, C.P.; Aguayo, A.; Aguilar-Hernandez, M.M. Analysis of RNA yield in extracellular vesicles isolated by membrane affinity column and differential ultracentrifugation. PLoS ONE 2020, 15, e0238545. [Google Scholar] [CrossRef]
  31. Nieuwland, R.; Siljander, P.R.; Falcon-Perez, J.M.; Witwer, K.W. Reproducibility of extracellular vesicle research. Eur. J. Cell Biol. 2022, 101, 151226. [Google Scholar] [CrossRef] [PubMed]
  32. Davidson, S.M.; Boulanger, C.M.; Aikawa, E.; Badimon, L.; Barile, L.; Binder, C.J.; Brisson, A.; Buzas, E.; Emanueli, C.; Jansen, F.; et al. Methods for the identification and characterization of extracellular vesicles in cardiovascular studies—From exosomes to microvesicles. Cardiovasc. Res. 2022. [Google Scholar] [CrossRef] [PubMed]
  33. Coly, P.M.; Boulanger, C.M. Role of extracellular vesicles in atherosclerosis: An update. J. Leukoc. Biol. 2022, 111, 51–62. [Google Scholar] [CrossRef] [PubMed]
  34. Bobryshev, Y.V.; Killingsworth, M.C.; Orekhov, A.N. Increased shedding of microvesicles from intimal smooth muscle cells in athero-prone areas of the human aorta: Implications for understanding of the predisease stage. Pathobiology 2013, 80, 24–31. [Google Scholar] [CrossRef]
  35. Leroyer, A.S.; Isobe, H.; Leseche, G.; Castier, Y.; Wassef, M.; Mallat, Z.; Binder, B.R.; Tedgui, A.; Boulanger, C.M. Cellular origins and thrombogenic activity of microparticles isolated from human atherosclerotic plaques. J. Am. Coll Cardiol. 2007, 49, 772–777. [Google Scholar] [CrossRef] [Green Version]
  36. Mayr, M.; Grainger, D.; Mayr, U.; Leroyer, A.S.; Leseche, G.; Sidibe, A.; Herbin, O.; Yin, X.; Gomes, A.; Madhu, B.; et al. Proteomics, metabolomics, and immunomics on microparticles derived from human atherosclerotic plaques. Circ. Cardiovasc Genet. 2009, 2, 379–388. [Google Scholar] [CrossRef] [Green Version]
  37. Canault, M.; Leroyer, A.S.; Peiretti, F.; Leseche, G.; Tedgui, A.; Bonardo, B.; Alessi, M.C.; Boulanger, C.M.; Nalbone, G. Microparticles of human atherosclerotic plaques enhance the shedding of the tumor necrosis factor-alpha converting enzyme/ADAM17 substrates, tumor necrosis factor and tumor necrosis factor receptor-1. Am. J. Pathol. 2007, 171, 1713–1723. [Google Scholar] [CrossRef] [Green Version]
  38. Alexander, Y.; Osto, E.; Schmidt-Trucksass, A.; Shechter, M.; Trifunovic, D.; Duncker, D.J.; Aboyans, V.; Back, M.; Badimon, L.; Cosentino, F.; et al. Endothelial function in cardiovascular medicine: A consensus paper of the European Society of Cardiology Working Groups on Atherosclerosis and Vascular Biology, Aorta and Peripheral Vascular Diseases, Coronary Pathophysiology and Microcirculation, and Thrombosis. Cardiovasc. Res. 2021, 117, 29–42. [Google Scholar] [CrossRef] [Green Version]
  39. Amabile, N.; Cheng, S.; Renard, J.M.; Larson, M.G.; Ghorbani, A.; McCabe, E.; Griffin, G.; Guerin, C.; Ho, J.E.; Shaw, S.Y.; et al. Association of circulating endothelial microparticles with cardiometabolic risk factors in the Framingham Heart Study. Eur. Heart J. 2014, 35, 2972–2979. [Google Scholar] [CrossRef]
  40. Diamant, M.; Nieuwland, R.; Pablo, R.F.; Sturk, A.; Smit, J.W.; Radder, J.K. Elevated numbers of tissue-factor exposing microparticles correlate with components of the metabolic syndrome in uncomplicated type 2 diabetes mellitus. Circulation 2002, 106, 2442–2447. [Google Scholar] [CrossRef] [Green Version]
  41. Ferreira, A.C.; Peter, A.A.; Mendez, A.J.; Jimenez, J.J.; Mauro, L.M.; Chirinos, J.A.; Ghany, R.; Virani, S.; Garcia, S.; Horstman, L.L.; et al. Postprandial hypertriglyceridemia increases circulating levels of endothelial cell microparticles. Circulation 2004, 110, 3599–3603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Heiss, C.; Amabile, N.; Lee, A.C.; Real, W.M.; Schick, S.F.; Lao, D.; Wong, M.L.; Jahn, S.; Angeli, F.S.; Minasi, P.; et al. Brief secondhand smoke exposure depresses endothelial progenitor cells activity and endothelial function: Sustained vascular injury and blunted nitric oxide production. J. Am. Coll Cardiol. 2008, 51, 1760–1771. [Google Scholar] [CrossRef] [PubMed]
  43. Koga, H.; Sugiyama, S.; Kugiyama, K.; Watanabe, K.; Fukushima, H.; Tanaka, T.; Sakamoto, T.; Yoshimura, M.; Jinnouchi, H.; Ogawa, H. Elevated levels of VE-cadherin-positive endothelial microparticles in patients with type 2 diabetes mellitus and coronary artery disease. J. Am. Coll Cardiol. 2005, 45, 1622–1630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Murugesan, S.; Hussey, H.; Saravanakumar, L.; Sinkey, R.G.; Sturdivant, A.B.; Powell, M.F.; Berkowitz, D.E. Extracellular Vesicles From Women With Severe Preeclampsia Impair Vascular Endothelial Function. Anesth. Analg. 2022, 134, 713–723. [Google Scholar] [CrossRef]
  45. Nomura, S.; Shouzu, A.; Omoto, S.; Nishikawa, M.; Fukuhara, S.; Iwasaka, T. Losartan and simvastatin inhibit platelet activation in hypertensive patients. J. Thromb. Thrombolysis 2004, 18, 177–185. [Google Scholar] [CrossRef]
  46. Ousmaal, M.E.F.; Gaceb, A.; Khene, M.A.; Ainouz, L.; Giaimis, J.; Andriantsitohaina, R.; Martinez, M.C.; Baz, A. Circulating microparticles released during dyslipidemia may exert deleterious effects on blood vessels and endothelial function. J. Diabetes Complicat. 2020, 34, 107683. [Google Scholar] [CrossRef]
  47. Preston, R.A.; Jy, W.; Jimenez, J.J.; Mauro, L.M.; Horstman, L.L.; Valle, M.; Aime, G.; Ahn, Y.S. Effects of severe hypertension on endothelial and platelet microparticles. Hypertension 2003, 41, 211–217. [Google Scholar] [CrossRef] [Green Version]
  48. Sabatier, F.; Darmon, P.; Hugel, B.; Combes, V.; Sanmarco, M.; Velut, J.G.; Arnoux, D.; Charpiot, P.; Freyssinet, J.M.; Oliver, C.; et al. Type 1 and type 2 diabetic patients display different patterns of cellular microparticles. Diabetes 2002, 51, 2840–2845. [Google Scholar] [CrossRef] [Green Version]
  49. Stockelman, K.A.; Hijmans, J.G.; Bammert, T.D.; Greiner, J.J.; Stauffer, B.L.; DeSouza, C.A. Circulating endothelial cell derived microvesicles are elevated with hypertension and associated with endothelial dysfunction. Can. J. Physiol. Pharmacol. 2020, 98, 557–561. [Google Scholar] [CrossRef]
  50. Suades, R.; Padro, T.; Alonso, R.; Lopez-Miranda, J.; Mata, P.; Badimon, L. Circulating CD45+/CD3+ lymphocyte-derived microparticles map lipid-rich atherosclerotic plaques in familial hypercholesterolaemia patients. Thromb. Haemost. 2014, 111, 111–121. [Google Scholar] [CrossRef]
  51. Cesaroni, G.; Forastiere, F.; Stafoggia, M.; Andersen, Z.J.; Badaloni, C.; Beelen, R.; Caracciolo, B.; de Faire, U.; Erbel, R.; Eriksen, K.T.; et al. Long term exposure to ambient air pollution and incidence of acute coronary events: Prospective cohort study and meta-analysis in 11 European cohorts from the ESCAPE Project. BMJ 2014, 348, f7412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Gao, Y.; Zhang, Q.; Sun, J.; Liang, Y.; Zhang, M.; Zhao, M.; Zhang, K.; Dong, C.; Ma, Q.; Liu, W.; et al. Extracellular vesicles derived from PM2.5-exposed alveolar epithelial cells mediate endothelial adhesion and atherosclerosis in ApoE(-/-) mice. FASEB J. 2022, 36, e22161. [Google Scholar] [CrossRef] [PubMed]
  53. Alharbi, M.G.; Lee, S.H.; Abdelazim, A.M.; Saadeldin, I.M.; Abomughaid, M.M. Role of Extracellular Vesicles in Compromising Cellular Resilience to Environmental Stressors. Biomed. Res. Int 2021, 2021, 9912281. [Google Scholar] [CrossRef]
  54. Rota, F.; Ferrari, L.; Hoxha, M.; Favero, C.; Antonioli, R.; Pergoli, L.; Greco, M.F.; Mariani, J.; Lazzari, L.; Bollati, V. Blood-derived extracellular vesicles isolated from healthy donors exposed to air pollution modulate in vitro endothelial cells behavior. Sci. Rep. 2020, 10, 20138. [Google Scholar] [CrossRef] [PubMed]
  55. Agouni, A.; Lagrue-Lak-Hal, A.H.; Ducluzeau, P.H.; Mostefai, H.A.; Draunet-Busson, C.; Leftheriotis, G.; Heymes, C.; Martinez, M.C.; Andriantsitohaina, R. Endothelial dysfunction caused by circulating microparticles from patients with metabolic syndrome. Am. J. Pathol. 2008, 173, 1210–1219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Boulanger, C.M.; Scoazec, A.; Ebrahimian, T.; Henry, P.; Mathieu, E.; Tedgui, A.; Mallat, Z. Circulating microparticles from patients with myocardial infarction cause endothelial dysfunction. Circulation 2001, 104, 2649–2652. [Google Scholar] [CrossRef]
  57. Brodsky, S.V.; Zhang, F.; Nasjletti, A.; Goligorsky, M.S. Endothelium-derived microparticles impair endothelial function in vitro. Am. J. Physiol. Heart Circ. Physiol. 2004, 286, H1910–H1915. [Google Scholar] [CrossRef] [Green Version]
  58. Jansen, F.; Yang, X.; Franklin, B.S.; Hoelscher, M.; Schmitz, T.; Bedorf, J.; Nickenig, G.; Werner, N. High glucose condition increases NADPH oxidase activity in endothelial microparticles that promote vascular inflammation. Cardiovasc. Res. 2013, 98, 94–106. [Google Scholar] [CrossRef] [Green Version]
  59. Martin, S.; Tesse, A.; Hugel, B.; Martinez, M.C.; Morel, O.; Freyssinet, J.M.; Andriantsitohaina, R. Shed membrane particles from T lymphocytes impair endothelial function and regulate endothelial protein expression. Circulation 2004, 109, 1653–1659. [Google Scholar] [CrossRef] [Green Version]
  60. Amabile, N.; Guerin, A.P.; Leroyer, A.; Mallat, Z.; Nguyen, C.; Boddaert, J.; London, G.M.; Tedgui, A.; Boulanger, C.M. Circulating endothelial microparticles are associated with vascular dysfunction in patients with end-stage renal failure. J. Am. Soc. Nephrol. 2005, 16, 3381–3388. [Google Scholar] [CrossRef] [Green Version]
  61. Osman, A.; El-Gamal, H.; Pasha, M.; Zeidan, A.; Korashy, H.M.; Abdelsalam, S.S.; Hasan, M.; Benameur, T.; Agouni, A. Endoplasmic Reticulum (ER) Stress-Generated Extracellular Vesicles (Microparticles) Self-Perpetuate ER Stress and Mediate Endothelial Cell Dysfunction Independently of Cell Survival. Front. Cardiovasc. Med. 2020, 7, 584791. [Google Scholar] [CrossRef] [PubMed]
  62. El Habhab, A.; Altamimy, R.; Abbas, M.; Kassem, M.; Amoura, L.; Qureshi, A.W.; El Itawi, H.; Kreutter, G.; Khemais-Benkhiat, S.; Zobairi, F.; et al. Significance of neutrophil microparticles in ischaemia-reperfusion: Pro-inflammatory effectors of endothelial senescence and vascular dysfunction. J. Cell Mol. Med. 2020, 24, 7266–7281. [Google Scholar] [CrossRef] [PubMed]
  63. Donadee, C.; Raat, N.J.; Kanias, T.; Tejero, J.; Lee, J.S.; Kelley, E.E.; Zhao, X.; Liu, C.; Reynolds, H.; Azarov, I.; et al. Nitric oxide scavenging by red blood cell microparticles and cell-free hemoglobin as a mechanism for the red cell storage lesion. Circulation 2011, 124, 465–476. [Google Scholar] [CrossRef] [PubMed]
  64. Peng, L.; Li, Y.; Li, X.; Du, Y.; Li, L.; Hu, C.; Zhang, J.; Qin, Y.; Wei, Y.; Zhang, H. Extracellular Vesicles Derived from Intermittent Hypoxia-Treated Red Blood Cells Impair Endothelial Function Through Regulating eNOS Phosphorylation and ET-1 Expression. Cardiovasc. Drugs Ther. 2021, 35, 901–913. [Google Scholar] [CrossRef]
  65. Brewster, L.M.; Bain, A.R.; Garcia, V.P.; Fandl, H.K.; Stone, R.; DeSouza, N.M.; Greiner, J.J.; Tymko, M.M.; Vizcardo-Galindo, G.A.; Figueroa-Mujica, R.J.; et al. Global REACH 2018: Dysfunctional extracellular microvesicles in Andean highlander males with excessive erythrocytosis. Am. J. Physiol. Heart Circ. Physiol. 2021, 320, H1851–H1861. [Google Scholar] [CrossRef]
  66. Taguchi, K.; Kaneko, N.; Okudaira, K.; Matsumoto, T.; Kobayashi, T. Endothelial dysfunction caused by circulating microparticles from diabetic mice is reduced by PD98059 through ERK and ICAM-1. Eur. J. Pharmacol. 2021, 913, 174630. [Google Scholar] [CrossRef]
  67. Abbas, M.; Jesel, L.; Auger, C.; Amoura, L.; Messas, N.; Manin, G.; Rumig, C.; Leon-Gonzalez, A.J.; Ribeiro, T.P.; Silva, G.C.; et al. Endothelial Microparticles From Acute Coronary Syndrome Patients Induce Premature Coronary Artery Endothelial Cell Aging and Thrombogenicity: Role of the Ang II/AT1 Receptor/NADPH Oxidase-Mediated Activation of MAPKs and PI3-Kinase Pathways. Circulation 2017, 135, 280–296. [Google Scholar] [CrossRef]
  68. Chatterjee, V.; Yang, X.; Ma, Y.; Cha, B.; Meegan, J.E.; Wu, M.; Yuan, S.Y. Endothelial microvesicles carrying Src-rich cargo impair adherens junction integrity and cytoskeleton homeostasis. Cardiovasc. Res. 2020, 116, 1525–1538. [Google Scholar] [CrossRef] [Green Version]
  69. Densmore, J.C.; Signorino, P.R.; Ou, J.; Hatoum, O.A.; Rowe, J.J.; Shi, Y.; Kaul, S.; Jones, D.W.; Sabina, R.E.; Pritchard, K.A., Jr.; et al. Endothelium-derived microparticles induce endothelial dysfunction and acute lung injury. Shock 2006, 26, 464–471. [Google Scholar] [CrossRef]
  70. Huber, L.C.; Jungel, A.; Distler, J.H.; Moritz, F.; Gay, R.E.; Michel, B.A.; Pisetsky, D.S.; Gay, S.; Distler, O. The role of membrane lipids in the induction of macrophage apoptosis by microparticles. Apoptosis 2007, 12, 363–374. [Google Scholar] [CrossRef] [Green Version]
  71. Abid Hussein, M.N.; Boing, A.N.; Sturk, A.; Hau, C.M.; Nieuwland, R. Inhibition of microparticle release triggers endothelial cell apoptosis and detachment. Thromb. Haemost. 2007, 98, 1096–1107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Boing, A.N.; Hau, C.M.; Sturk, A.; Nieuwland, R. Platelet microparticles contain active caspase 3. Platelets 2008, 19, 96–103. [Google Scholar] [CrossRef] [PubMed]
  73. Sarkar, A.; Mitra, S.; Mehta, S.; Raices, R.; Wewers, M.D. Monocyte derived microvesicles deliver a cell death message via encapsulated caspase-1. PLoS ONE 2009, 4, e7140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Abid Hussein, M.N.; Nieuwland, R.; Hau, C.M.; Evers, L.M.; Meesters, E.W.; Sturk, A. Cell-derived microparticles contain caspase 3 in vitro and in vivo. J. Thromb. Haemost. 2005, 3, 888–896. [Google Scholar] [CrossRef]
  75. Jansen, F.; Yang, X.; Hoyer, F.F.; Paul, K.; Heiermann, N.; Becher, M.U.; Abu Hussein, N.; Kebschull, M.; Bedorf, J.; Franklin, B.S.; et al. Endothelial microparticle uptake in target cells is annexin I/phosphatidylserine receptor dependent and prevents apoptosis. Arterioscler. Thromb. Vasc. Biol. 2012, 32, 1925–1935. [Google Scholar] [CrossRef] [Green Version]
  76. Li, L.; Wang, H.; Zhang, J.; Chen, X.; Zhang, Z.; Li, Q. Effect of endothelial progenitor cell-derived extracellular vesicles on endothelial cell ferroptosis and atherosclerotic vascular endothelial injury. Cell Death Discov. 2021, 7, 235. [Google Scholar] [CrossRef]
  77. Bao, H.; Chen, Y.X.; Huang, K.; Zhuang, F.; Bao, M.; Han, Y.; Chen, X.H.; Shi, Q.; Yao, Q.P.; Qi, Y.X. Platelet-derived microparticles promote endothelial cell proliferation in hypertension via miR-142-3p. FASEB J. 2018, 32, 3912–3923. [Google Scholar] [CrossRef] [Green Version]
  78. Zu, L.; Ren, C.; Pan, B.; Zhou, B.; Zhou, E.; Niu, C.; Wang, X.; Zhao, M.; Gao, W.; Guo, L.; et al. Endothelial microparticles after antihypertensive and lipid-lowering therapy inhibit the adhesion of monocytes to endothelial cells. Int. J. Cardiol. 2016, 202, 756–759. [Google Scholar] [CrossRef]
  79. Guerrero, F.; Carmona, A.; Obrero, T.; Jimenez, M.J.; Soriano, S.; Moreno, J.A.; Martin-Malo, A.; Aljama, P. Role of endothelial microvesicles released by p-cresol on endothelial dysfunction. Sci. Rep. 2020, 10, 10657. [Google Scholar] [CrossRef]
  80. Sanz-Rubio, D.; Khalyfa, A.; Qiao, Z.; Ullate, J.; Marin, J.M.; Kheirandish-Gozal, L.; Gozal, D. Cell-Selective Altered Cargo Properties of Extracellular Vesicles Following In Vitro Exposures to Intermittent Hypoxia. Int. J. Mol. Sci. 2021, 22, 5604. [Google Scholar] [CrossRef]
  81. Li, Y.; Zhang, H.; Du, Y.; Peng, L.; Qin, Y.; Liu, H.; Ma, X.; Wei, Y. Extracellular vesicle microRNA cargoes from intermittent hypoxia-exposed cardiomyocytes and their effect on endothelium. Biochem. Biophys. Res. Commun. 2021, 548, 182–188. [Google Scholar] [CrossRef] [PubMed]
  82. Gao, W.; Guo, X.; Wang, Y.; Jian, D.; Li, M. Monocyte-derived extracellular vesicles upon treated by palmitate promote endothelial migration and monocytes attachment to endothelial cells. Biochem. Biophys. Res. Commun. 2020, 523, 685–691. [Google Scholar] [CrossRef] [PubMed]
  83. Ajikumar, A.; Long, M.B.; Heath, P.R.; Wharton, S.B.; Ince, P.G.; Ridger, V.C.; Simpson, J.E. Neutrophil-Derived Microvesicle Induced Dysfunction of Brain Microvascular Endothelial Cells In Vitro. Int. J. Mol. Sci. 2019, 20, 5227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Bjorkegren, J.L.M.; Lusis, A.J. Atherosclerosis: Recent developments. Cell 2022, 185, 1630–1645. [Google Scholar] [CrossRef] [PubMed]
  85. Mesri, M.; Altieri, D.C. Endothelial cell activation by leukocyte microparticles. J. Immunol. 1998, 161, 4382–4387. [Google Scholar] [PubMed]
  86. Mesri, M.; Altieri, D.C. Leukocyte microparticles stimulate endothelial cell cytokine release and tissue factor induction in a JNK1 signaling pathway. J. Biol. Chem. 1999, 274, 23111–23118. [Google Scholar] [CrossRef] [Green Version]
  87. Robbins, P.D.; Dorronsoro, A.; Booker, C.N. Regulation of chronic inflammatory and immune processes by extracellular vesicles. J. Clin. Investig. 2016, 126, 1173–1180. [Google Scholar] [CrossRef] [Green Version]
  88. Burger, D.; Turner, M.; Xiao, F.; Munkonda, M.N.; Akbari, S.; Burns, K.D. High glucose increases the formation and pro-oxidative activity of endothelial microparticles. Diabetologia 2017, 60, 1791–1800. [Google Scholar] [CrossRef] [Green Version]
  89. Barry, O.P.; Pratico, D.; Savani, R.C.; FitzGerald, G.A. Modulation of monocyte-endothelial cell interactions by platelet microparticles. J. Clin. Investig. 1998, 102, 136–144. [Google Scholar] [CrossRef]
  90. Nomura, S.; Tandon, N.N.; Nakamura, T.; Cone, J.; Fukuhara, S.; Kambayashi, J. High-shear-stress-induced activation of platelets and microparticles enhances expression of cell adhesion molecules in THP-1 and endothelial cells. Atherosclerosis 2001, 158, 277–287. [Google Scholar] [CrossRef]
  91. Forlow, S.B.; McEver, R.P.; Nollert, M.U. Leukocyte-leukocyte interactions mediated by platelet microparticles under flow. Blood 2000, 95, 1317–1323. [Google Scholar] [CrossRef] [PubMed]
  92. Mause, S.F.; von Hundelshausen, P.; Zernecke, A.; Koenen, R.R.; Weber, C. Platelet microparticles: A transcellular delivery system for RANTES promoting monocyte recruitment on endothelium. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 1512–1518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Vasina, E.M.; Cauwenberghs, S.; Feijge, M.A.; Heemskerk, J.W.; Weber, C.; Koenen, R.R. Microparticles from apoptotic platelets promote resident macrophage differentiation. Cell Death Dis. 2011, 2, e211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Chimen, M.; Evryviadou, A.; Box, C.L.; Harrison, M.J.; Hazeldine, J.; Dib, L.H.; Kuravi, S.J.; Payne, H.; Price, J.M.J.; Kavanagh, D.; et al. Appropriation of GPIbalpha from platelet-derived extracellular vesicles supports monocyte recruitment in systemic inflammation. Haematologica 2020, 105, 1248–1261. [Google Scholar] [CrossRef]
  95. Mause, S.F.; Ritzel, E.; Liehn, E.A.; Hristov, M.; Bidzhekov, K.; Muller-Newen, G.; Soehnlein, O.; Weber, C. Platelet microparticles enhance the vasoregenerative potential of angiogenic early outgrowth cells after vascular injury. Circulation 2010, 122, 495–506. [Google Scholar] [CrossRef] [Green Version]
  96. Wadey, R.M.; Connolly, K.D.; Mathew, D.; Walters, G.; Rees, D.A.; James, P.E. Inflammatory adipocyte-derived extracellular vesicles promote leukocyte attachment to vascular endothelial cells. Atherosclerosis 2019, 283, 19–27. [Google Scholar] [CrossRef] [Green Version]
  97. Pitanga, T.N.; de Aragao Franca, L.; Rocha, V.C.; Meirelles, T.; Borges, V.M.; Goncalves, M.S.; Pontes-de-Carvalho, L.C.; Noronha-Dutra, A.A.; dos-Santos, W.L. Neutrophil-derived microparticles induce myeloperoxidase-mediated damage of vascular endothelial cells. BMC Cell Biol. 2014, 15, 21. [Google Scholar] [CrossRef] [Green Version]
  98. Rossaint, J.; Kuhne, K.; Skupski, J.; Van Aken, H.; Looney, M.R.; Hidalgo, A.; Zarbock, A. Directed transport of neutrophil-derived extracellular vesicles enables platelet-mediated innate immune response. Nat. Commun. 2016, 7, 13464. [Google Scholar] [CrossRef]
  99. Gomez, I.; Ward, B.; Souilhol, C.; Recarti, C.; Ariaans, M.; Johnston, J.; Burnett, A.; Mahmoud, M.; Luong, L.A.; West, L.; et al. Neutrophil microvesicles drive atherosclerosis by delivering miR-155 to atheroprone endothelium. Nat. Commun. 2020, 11, 214. [Google Scholar] [CrossRef] [Green Version]
  100. Tang, N.; Sun, B.; Gupta, A.; Rempel, H.; Pulliam, L. Monocyte exosomes induce adhesion molecules and cytokines via activation of NF-kappaB in endothelial cells. FASEB J. 2016, 30, 3097–3106. [Google Scholar] [CrossRef] [Green Version]
  101. Wang, J.G.; Williams, J.C.; Davis, B.K.; Jacobson, K.; Doerschuk, C.M.; Ting, J.P.; Mackman, N. Monocytic microparticles activate endothelial cells in an IL-1beta-dependent manner. Blood 2011, 118, 2366–2374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Mills, E.L.; Kelly, B.; Logan, A.; Costa, A.S.H.; Varma, M.; Bryant, C.E.; Tourlomousis, P.; Dabritz, J.H.M.; Gottlieb, E.; Latorre, I.; et al. Succinate Dehydrogenase Supports Metabolic Repurposing of Mitochondria to Drive Inflammatory Macrophages. Cell 2016, 167, 457–470 e413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. West, A.P.; Shadel, G.S. Mitochondrial DNA in innate immune responses and inflammatory pathology. Nat. Rev. Immunol. 2017, 17, 363–375. [Google Scholar] [CrossRef] [PubMed]
  104. 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]
  105. Zhang, Q.; Raoof, M.; Chen, Y.; Sumi, Y.; Sursal, T.; Junger, W.; Brohi, K.; Itagaki, K.; Hauser, C.J. Circulating mitochondrial DAMPs cause inflammatory responses to injury. Nature 2010, 464, 104–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. 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] [Green Version]
  107. Soubannier, V.; Rippstein, P.; Kaufman, B.A.; Shoubridge, E.A.; McBride, H.M. Reconstitution of mitochondria derived vesicle formation demonstrates selective enrichment of oxidized cargo. PLoS ONE 2012, 7, e52830. [Google Scholar] [CrossRef] [Green Version]
  108. 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]
  109. Holvoet, P.; Vanhaverbeke, M.; Bloch, K.; Baatsen, P.; Sinnaeve, P.; Janssens, S. Low MT-CO1 in Monocytes and Microvesicles Is Associated With Outcome in Patients With Coronary Artery Disease. J. Am. Heart Assoc. 2016, 5, e004207. [Google Scholar] [CrossRef]
  110. Boudreau, L.H.; Duchez, A.C.; Cloutier, N.; Soulet, D.; Martin, N.; Bollinger, J.; Pare, A.; Rousseau, M.; Naika, G.S.; Levesque, 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] [Green Version]
  111. He, S.; Wu, C.; Xiao, J.; Li, D.; Sun, Z.; Li, M. Endothelial extracellular vesicles modulate the macrophage phenotype: Potential implications in atherosclerosis. Scand. J. Immunol. 2018, 87, e12648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Liu, Y.; Li, Q.; Hosen, M.R.; Zietzer, A.; Flender, A.; Levermann, P.; Schmitz, T.; Fruhwald, D.; Goody, P.; Nickenig, G.; et al. Atherosclerotic Conditions Promote the Packaging of Functional MicroRNA-92a-3p Into Endothelial Microvesicles. Circ. Res. 2019, 124, 575–587. [Google Scholar] [CrossRef] [PubMed]
  113. Yang, M.; Walker, S.A.; Aguilar Diaz de Leon, J.S.; Davidovich, I.; Broad, K.; Talmon, Y.; Borges, C.R.; Wolfram, J. Extracellular vesicle glucose transporter-1 and glycan features in monocyte-endothelial inflammatory interactions. Nanomedicine 2022, 42, 102515. [Google Scholar] [CrossRef]
  114. Perez-Casal, M.; Downey, C.; Cutillas-Moreno, B.; Zuzel, M.; Fukudome, K.; Toh, C.H. Microparticle-associated endothelial protein C receptor and the induction of cytoprotective and anti-inflammatory effects. Haematologica 2009, 94, 387–394. [Google Scholar] [CrossRef] [Green Version]
  115. Luquero, A.; Vilahur, G.; Crespo, J.; Badimon, L.; Borrell-Pages, M. Microvesicles carrying LRP5 induce macrophage polarization to an anti-inflammatory phenotype. J. Cell Mol. Med. 2021, 25, 7935–7947. [Google Scholar] [CrossRef] [PubMed]
  116. Hergenreider, E.; Heydt, S.; Treguer, 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]
  117. Gasser, O.; Schifferli, J.A. Activated polymorphonuclear neutrophils disseminate anti-inflammatory microparticles by ectocytosis. Blood 2004, 104, 2543–2548. [Google Scholar] [CrossRef]
  118. Dalli, J.; Norling, L.V.; Renshaw, D.; Cooper, D.; Leung, K.Y.; Perretti, M. Annexin 1 mediates the rapid anti-inflammatory effects of neutrophil-derived microparticles. Blood 2008, 112, 2512–2519. [Google Scholar] [CrossRef] [Green Version]
  119. Jansen, F.; Yang, X.; Baumann, K.; Przybilla, D.; Schmitz, T.; Flender, A.; Paul, K.; Alhusseiny, A.; Nickenig, G.; Werner, N. Endothelial microparticles reduce ICAM-1 expression in a microRNA-222-dependent mechanism. J. Cell Mol. Med. 2015, 19, 2202–2214. [Google Scholar] [CrossRef]
  120. 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] [Green Version]
  121. Koppler, B.; Cohen, C.; Schlondorff, D.; Mack, M. Differential mechanisms of microparticle transfer toB cells and monocytes: Anti-inflammatory propertiesof microparticles. Eur. J. Immunol. 2006, 36, 648–660. [Google Scholar] [CrossRef] [PubMed]
  122. Hosseinkhani, B.; Kuypers, S.; van den Akker, N.M.S.; Molin, D.G.M.; Michiels, L. Extracellular Vesicles Work as a Functional Inflammatory Mediator Between Vascular Endothelial Cells and Immune Cells. Front. Immunol. 2018, 9, 1789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Hosseinkhani, B.; van den Akker, N.M.S.; Molin, D.G.M.; Michiels, L. (Sub)populations of extracellular vesicles released by TNF-alpha -triggered human endothelial cells promote vascular inflammation and monocyte migration. J. Extracell Vesicles 2020, 9, 1801153. [Google Scholar] [CrossRef] [PubMed]
  124. Edrissi, H.; Schock, S.C.; Hakim, A.M.; Thompson, C.S. Microparticles generated during chronic cerebral ischemia increase the permeability of microvascular endothelial barriers in vitro. Brain Res. 2016, 1634, 83–93. [Google Scholar] [CrossRef]
  125. Marcos-Ramiro, B.; Oliva Nacarino, P.; Serrano-Pertierra, E.; Blanco-Gelaz, M.A.; Weksler, B.B.; Romero, I.A.; Couraud, P.O.; Tunon, A.; Lopez-Larrea, C.; Millan, J.; et al. Microparticles in multiple sclerosis and clinically isolated syndrome: Effect on endothelial barrier function. BMC Neurosci. 2014, 15, 110. [Google Scholar] [CrossRef] [Green Version]
  126. Lapping-Carr, G.; Gemel, J.; Mao, Y.; Sparks, G.; Harrington, M.; Peddinti, R.; Beyer, E.C. Insights image for “Circulating extracellular vesicles from patients with acute chest syndrome disrupt adherens junctions between endothelial cells”. Pediatr. Res. 2021, 89, 1036. [Google Scholar] [CrossRef]
  127. Huber, J.; Vales, A.; Mitulovic, G.; Blumer, M.; Schmid, R.; Witztum, J.L.; Binder, B.R.; Leitinger, N. Oxidized membrane vesicles and blebs from apoptotic cells contain biologically active oxidized phospholipids that induce monocyte-endothelial interactions. Arterioscler. Thromb. Vasc. Biol. 2002, 22, 101–107. [Google Scholar] [CrossRef] [Green Version]
  128. Fink, K.; Moebes, M.; Vetter, C.; Bourgeois, N.; Schmid, B.; Bode, C.; Helbing, T.; Busch, H.J. Selenium prevents microparticle-induced endothelial inflammation in patients after cardiopulmonary resuscitation. Crit. Care 2015, 19, 58. [Google Scholar] [CrossRef] [Green Version]
  129. Wang, Y.; Han, B.; Wang, Y.; Wang, C.; Zhang, H.; Xue, J.; Wang, X.; Niu, T.; Niu, Z.; Chen, Y. Mesenchymal stem cell-secreted extracellular vesicles carrying TGF-beta1 up-regulate miR-132 and promote mouse M2 macrophage polarization. J. Cell Mol. Med. 2020, 24, 12750–12764. [Google Scholar] [CrossRef]
  130. Rautou, P.E.; Leroyer, A.S.; Ramkhelawon, B.; Devue, C.; Duflaut, D.; Vion, A.C.; Nalbone, G.; Castier, Y.; Leseche, G.; Lehoux, S.; et al. Microparticles from human atherosclerotic plaques promote endothelial ICAM-1-dependent monocyte adhesion and transendothelial migration. Circ. Res. 2011, 108, 335–343. [Google Scholar] [CrossRef] [Green Version]
  131. Sanchez-Rodriguez, M.B.; Tellez, E.; Casulleras, M.; Borras, F.E.; Arroyo, V.; Claria, J.; Sarrias, M.R. Reduced Plasma Extracellular Vesicle CD5L Content in Patients With Acute-On-Chronic Liver Failure: Interplay With Specialized Pro-Resolving Lipid Mediators. Front. Immunol. 2022, 13, 842996. [Google Scholar] [CrossRef] [PubMed]
  132. Fitzsimons, S.; Oggero, S.; Bruen, R.; McCarthy, C.; Strowitzki, M.J.; Mahon, N.G.; Ryan, N.; Brennan, E.P.; Barry, M.; Perretti, M.; et al. microRNA-155 Is Decreased During Atherosclerosis Regression and Is Increased in Urinary Extracellular Vesicles During Atherosclerosis Progression. Front. Immunol. 2020, 11, 576516. [Google Scholar] [CrossRef] [PubMed]
  133. Nguyen, M.A.; Karunakaran, D.; Geoffrion, M.; Cheng, H.S.; Tandoc, K.; Perisic Matic, L.; Hedin, U.; Maegdefessel, L.; Fish, J.E.; Rayner, K.J. Extracellular Vesicles Secreted by Atherogenic Macrophages Transfer MicroRNA to Inhibit Cell Migration. Arterioscler. Thromb. Vasc. Biol. 2018, 38, 49–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Liu, M.L.; Scalia, R.; Mehta, J.L.; Williams, K.J. Cholesterol-induced membrane microvesicles as novel carriers of damage-associated molecular patterns: Mechanisms of formation, action, and detoxification. Arterioscler. Thromb. Vasc. Biol. 2012, 32, 2113–2121. [Google Scholar] [CrossRef] [Green Version]
  135. Hoyer, F.F.; Giesen, M.K.; Nunes Franca, C.; Lutjohann, D.; Nickenig, G.; Werner, N. Monocytic microparticles promote atherogenesis by modulating inflammatory cells in mice. J. Cell Mol. Med. 2012, 16, 2777–2788. [Google Scholar] [CrossRef] [Green Version]
  136. Ismail, N.; Wang, Y.; Dakhlallah, D.; Moldovan, L.; Agarwal, K.; Batte, K.; Shah, P.; Wisler, J.; Eubank, T.D.; Tridandapani, S.; et al. Macrophage microvesicles induce macrophage differentiation and miR-223 transfer. Blood 2013, 121, 984–995. [Google Scholar] [CrossRef] [Green Version]
  137. Chang, Y.J.; Li, Y.S.; Wu, C.C.; Wang, K.C.; Huang, T.C.; Chen, Z.; Chien, S. Extracellular MicroRNA-92a Mediates Endothelial Cell-Macrophage Communication. Arterioscler. Thromb. Vasc. Biol. 2019, 39, 2492–2504. [Google Scholar] [CrossRef]
  138. Feng, C.; Chen, Q.; Fan, M.; Guo, J.; Liu, Y.; Ji, T.; Zhu, J.; Zhao, X. Platelet-derived microparticles promote phagocytosis of oxidized low-density lipoprotein by macrophages, potentially enhancing foam cell formation. Ann. Transl. Med. 2019, 7, 477. [Google Scholar] [CrossRef]
  139. Distler, J.H.; Huber, L.C.; Hueber, A.J.; Reich, C.F., 3rd; Gay, S.; Distler, O.; Pisetsky, D.S. The release of microparticles by apoptotic cells and their effects on macrophages. Apoptosis 2005, 10, 731–741. [Google Scholar] [CrossRef]
  140. Pizzirani, C.; Ferrari, D.; Chiozzi, P.; Adinolfi, E.; Sandona, D.; Savaglio, E.; Di Virgilio, F. Stimulation of P2 receptors causes release of IL-1beta-loaded microvesicles from human dendritic cells. Blood 2007, 109, 3856–3864. [Google Scholar] [CrossRef] [Green Version]
  141. Tsiantoulas, D.; Perkmann, T.; Afonyushkin, T.; Mangold, A.; Prohaska, T.A.; Papac-Milicevic, N.; Millischer, V.; Bartel, C.; Horkko, S.; Boulanger, C.M.; et al. Circulating microparticles carry oxidation-specific epitopes and are recognized by natural IgM antibodies. J. Lipid Res. 2015, 56, 440–448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Obermayer, G.; Afonyushkin, T.; Goderle, L.; Puhm, F.; Schrottmaier, W.; Taqi, S.; Schwameis, M.; Ay, C.; Pabinger, I.; Jilma, B.; et al. Natural IgM antibodies inhibit microvesicle-driven coagulation and thrombosis. Blood 2021, 137, 1406–1415. [Google Scholar] [CrossRef] [PubMed]
  143. Vajen, T.; Benedikter, B.J.; Heinzmann, A.C.A.; Vasina, E.M.; Henskens, Y.; Parsons, M.; Maguire, P.B.; Stassen, F.R.; Heemskerk, J.W.M.; Schurgers, L.J.; et al. Platelet extracellular vesicles induce a pro-inflammatory smooth muscle cell phenotype. J. Extracell Vesicles 2017, 6, 1322454. [Google Scholar] [CrossRef] [PubMed]
  144. Niu, C.; Wang, X.; Zhao, M.; Cai, T.; Liu, P.; Li, J.; Willard, B.; Zu, L.; Zhou, E.; Li, Y.; et al. Macrophage Foam Cell-Derived Extracellular Vesicles Promote Vascular Smooth Muscle Cell Migration and Adhesion. J. Am. Heart Assoc. 2016, 5, e004099. [Google Scholar] [CrossRef] [Green Version]
  145. Pakala, R.; Rha, S.W.; Kuchulakanti, P.K.; Cheneau, E.; Baffour, R.; Waksman, R. Peroxisome proliferator-activated receptor gamma; Its role in atherosclerosis and restenosis. Cardiovasc. Radiat. Med. 2004, 5, 44–48. [Google Scholar] [CrossRef]
  146. Weber, A.A.; Schror, K. The significance of platelet-derived growth factors for proliferation of vascular smooth muscle cells. Platelets 1999, 10, 77–96. [Google Scholar] [CrossRef]
  147. Li, S.S.; Gao, S.; Chen, Y.; Bao, H.; Li, Z.T.; Yao, Q.P.; Liu, J.T.; Wang, Y.; Qi, Y.X. Platelet-derived microvesicles induce calcium oscillations and promote VSMC migration via TRPV4. Theranostics 2021, 11, 2410–2423. [Google Scholar] [CrossRef]
  148. He, C.; Hu, X.; Weston, T.A.; Jung, R.S.; Sandhu, J.; Huang, S.; Heizer, P.; Kim, J.; Ellison, R.; Xu, J.; et al. Macrophages release plasma membrane-derived particles rich in accessible cholesterol. Proc. Natl. Acad. Sci. USA 2018, 115, E8499–E8508. [Google Scholar] [CrossRef] [Green Version]
  149. Wang, Q.; Dong, Y.; Wang, H. microRNA-19b-3p-containing extracellular vesicles derived from macrophages promote the development of atherosclerosis by targeting JAZF1. J. Cell Mol. Med. 2022, 26, 48–59. [Google Scholar] [CrossRef]
  150. Brousseau, C.; Morissette, G.; Fortin, J.P.; Marceau, F.; Petitclerc, E. Tumor cells expressing tissue factor influence the migration of smooth muscle cells in a catalytic activity-dependent way. Can. J. Physiol. Pharmacol. 2009, 87, 694–701. [Google Scholar] [CrossRef]
  151. Paudel, K.R.; Oak, M.H.; Kim, D.W. Smooth Muscle Cell Derived Microparticles Acts as Autocrine Activation of Smooth Muscle Cell Proliferation by Mitogen Associated Protein Kinase Upregulation. J. Nanosci. Nanotechnol. 2020, 20, 5746–5750. [Google Scholar] [CrossRef] [PubMed]
  152. Perdomo, L.; Vidal-Gomez, X.; Soleti, R.; Vergori, L.; Duluc, L.; Chwastyniak, M.; Bisserier, M.; Le Lay, S.; Villard, A.; Simard, G.; et al. Large Extracellular Vesicle-Associated Rap1 Accumulates in Atherosclerotic Plaques, Correlates With Vascular Risks and Is Involved in Atherosclerosis. Circ. Res. 2020, 127, 747–760. [Google Scholar] [CrossRef] [PubMed]
  153. Gao, X.F.; Wang, Z.M.; Chen, A.Q.; Wang, F.; Luo, S.; Gu, Y.; Kong, X.Q.; Zuo, G.F.; Jiang, X.M.; Ding, G.W.; et al. Plasma Small Extracellular Vesicle-Carried miRNA-501-5p Promotes Vascular Smooth Muscle Cell Phenotypic Modulation-Mediated In-Stent Restenosis. Oxid. Med. Cell Longev. 2021, 2021, 6644970. [Google Scholar] [CrossRef] [PubMed]
  154. Shan, Z.; Qin, S.; Li, W.; Wu, W.; Yang, J.; Chu, M.; Li, X.; Huo, Y.; Schaer, G.L.; Wang, S.; et al. An Endocrine Genetic Signal Between Blood Cells and Vascular Smooth Muscle Cells: Role of MicroRNA-223 in Smooth Muscle Function and Atherogenesis. J. Am. Coll Cardiol. 2015, 65, 2526–2537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Wang, Y.; Xu, Z.; Wang, X.; Zheng, J.; Peng, L.; Zhou, Y.; Song, Y.; Lu, Z. Extracellular-vesicle containing miRNA-503-5p released by macrophages contributes to atherosclerosis. Aging 2021, 13, 12239–12257. [Google Scholar] [CrossRef]
  156. Liu, R.; Shen, H.; Ma, J.; Sun, L.; Wei, M. Extracellular Vesicles Derived from Adipose Mesenchymal Stem Cells Regulate the Phenotype of Smooth Muscle Cells to Limit Intimal Hyperplasia. Cardiovasc. Drugs Ther. 2016, 30, 111–118. [Google Scholar] [CrossRef]
  157. Mas-Bargues, C.; Borras, C.; Alique, M. The Contribution of Extracellular Vesicles From Senescent Endothelial and Vascular Smooth Muscle Cells to Vascular Calcification. Front. Cardiovasc. Med. 2022, 9, 854726. [Google Scholar] [CrossRef]
  158. Alique, M.; Ruiz-Torres, M.P.; Bodega, G.; Noci, M.V.; Troyano, N.; Bohorquez, L.; Luna, C.; Luque, R.; Carmona, A.; Carracedo, J.; et al. Microvesicles from the plasma of elderly subjects and from senescent endothelial cells promote vascular calcification. Aging 2017, 9, 778–789. [Google Scholar] [CrossRef] [Green Version]
  159. Schurgers, L.J.; Akbulut, A.C.; Kaczor, D.M.; Halder, M.; Koenen, R.R.; Kramann, R. Initiation and Propagation of Vascular Calcification Is Regulated by a Concert of Platelet- and Smooth Muscle Cell-Derived Extracellular Vesicles. Front. Cardiovasc. Med. 2018, 5, 36. [Google Scholar] [CrossRef]
  160. Hutcheson, J.D.; Goettsch, C.; Bertazzo, S.; Maldonado, N.; Ruiz, J.L.; Goh, W.; Yabusaki, K.; Faits, T.; Bouten, C.; Franck, G.; et al. Genesis and growth of extracellular-vesicle-derived microcalcification in atherosclerotic plaques. Nat. Mater. 2016, 15, 335–343. [Google Scholar] [CrossRef] [Green Version]
  161. New, S.E.; Goettsch, C.; Aikawa, M.; Marchini, J.F.; Shibasaki, M.; Yabusaki, K.; Libby, P.; Shanahan, C.M.; Croce, K.; Aikawa, E. Macrophage-derived matrix vesicles: An alternative novel mechanism for microcalcification in atherosclerotic plaques. Circ. Res. 2013, 113, 72–77. [Google Scholar] [CrossRef] [PubMed]
  162. New, S.E.; Aikawa, E. Role of extracellular vesicles in de novo mineralization: An additional novel mechanism of cardiovascular calcification. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 1753–1758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Reynolds, J.L.; Joannides, A.J.; Skepper, J.N.; McNair, R.; Schurgers, L.J.; Proudfoot, D.; Jahnen-Dechent, W.; Weissberg, P.L.; Shanahan, C.M. Human vascular smooth muscle cells undergo vesicle-mediated calcification in response to changes in extracellular calcium and phosphate concentrations: A potential mechanism for accelerated vascular calcification in ESRD. J. Am. Soc. Nephrol. 2004, 15, 2857–2867. [Google Scholar] [CrossRef] [Green Version]
  164. Buendia, P.; Montes de Oca, A.; Madueno, J.A.; Merino, A.; Martin-Malo, A.; Aljama, P.; Ramirez, R.; Rodriguez, M.; Carracedo, J. Endothelial microparticles mediate inflammation-induced vascular calcification. FASEB J. 2015, 29, 173–181. [Google Scholar] [CrossRef] [PubMed]
  165. Alique, M.; Bodega, G.; Corchete, E.; Garcia-Menendez, E.; de Sequera, P.; Luque, R.; Rodriguez-Padron, D.; Marques, M.; Portoles, J.; Carracedo, J.; et al. Microvesicles from indoxyl sulfate-treated endothelial cells induce vascular calcification in vitro. Comput. Struct. Biotechnol. J. 2020, 18, 953–966. [Google Scholar] [CrossRef] [PubMed]
  166. Goettsch, C.; Hutcheson, J.D.; Aikawa, M.; Iwata, H.; Pham, T.; Nykjaer, A.; Kjolby, M.; Rogers, M.; Michel, T.; Shibasaki, M.; et al. Sortilin mediates vascular calcification via its recruitment into extracellular vesicles. J. Clin. Investig. 2016, 126, 1323–1336. [Google Scholar] [CrossRef] [PubMed]
  167. Freise, C.; Querfeld, U.; Ludwig, A.; Hamm, B.; Schnorr, J.; Taupitz, M. Uraemic extracellular vesicles augment osteogenic transdifferentiation of vascular smooth muscle cells via enhanced AKT signalling and PiT-1 expression. J. Cell Mol. Med. 2021, 25, 5602–5614. [Google Scholar] [CrossRef]
  168. Furmanik, M.; Chatrou, M.; van Gorp, R.; Akbulut, A.; Willems, B.; Schmidt, H.; van Eys, G.; Bochaton-Piallat, M.L.; Proudfoot, D.; Biessen, E.; et al. Reactive Oxygen-Forming Nox5 Links Vascular Smooth Muscle Cell Phenotypic Switching and Extracellular Vesicle-Mediated Vascular Calcification. Circ. Res. 2020, 127, 911–927. [Google Scholar] [CrossRef]
  169. Mallat, Z.; Hugel, B.; Ohan, J.; Leseche, G.; Freyssinet, J.M.; Tedgui, A. Shed membrane microparticles with procoagulant potential in human atherosclerotic plaques: A role for apoptosis in plaque thrombogenicity. Circulation 1999, 99, 348–353. [Google Scholar] [CrossRef] [Green Version]
  170. Lawler, P.R.; Bhatt, D.L.; Godoy, L.C.; Luscher, T.F.; Bonow, R.O.; Verma, S.; Ridker, P.M. Targeting cardiovascular inflammation: Next steps in clinical translation. Eur. Heart J. 2021, 42, 113–131. [Google Scholar] [CrossRef]
  171. Stark, K.; Massberg, S. Interplay between inflammation and thrombosis in cardiovascular pathology. Nat. Rev. Cardiol. 2021, 18, 666–682. [Google Scholar] [CrossRef] [PubMed]
  172. Sedding, D.G.; Boyle, E.C.; Demandt, J.A.F.; Sluimer, J.C.; Dutzmann, J.; Haverich, A.; Bauersachs, J. Vasa Vasorum Angiogenesis: Key Player in the Initiation and Progression of Atherosclerosis and Potential Target for the Treatment of Cardiovascular Disease. Front. Immunol. 2018, 9, 706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Michel, J.B.; Virmani, R.; Arbustini, E.; Pasterkamp, G. Intraplaque haemorrhages as the trigger of plaque vulnerability. Eur. Heart J. 2011, 32, 1977–1985. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Soleti, R.; Benameur, T.; Porro, C.; Panaro, M.A.; Andriantsitohaina, R.; Martinez, M.C. Microparticles harboring Sonic Hedgehog promote angiogenesis through the upregulation of adhesion proteins and proangiogenic factors. Carcinogenesis 2009, 30, 580–588. [Google Scholar] [CrossRef]
  175. Mezentsev, A.; Merks, R.M.; O’Riordan, E.; Chen, J.; Mendelev, N.; Goligorsky, M.S.; Brodsky, S.V. Endothelial microparticles affect angiogenesis in vitro: Role of oxidative stress. Am. J. Physiol. Heart Circ. Physiol. 2005, 289, H1106–H1114. [Google Scholar] [CrossRef]
  176. Deregibus, M.C.; Cantaluppi, V.; Calogero, R.; Lo Iacono, M.; Tetta, C.; Biancone, L.; Bruno, S.; Bussolati, B.; Camussi, G. Endothelial progenitor cell derived microvesicles activate an angiogenic program in endothelial cells by a horizontal transfer of mRNA. Blood 2007, 110, 2440–2448. [Google Scholar] [CrossRef] [Green Version]
  177. Yamamoto, S.; Niida, S.; Azuma, E.; Yanagibashi, T.; Muramatsu, M.; Huang, T.T.; Sagara, H.; Higaki, S.; Ikutani, M.; Nagai, Y.; et al. Inflammation-induced endothelial cell-derived extracellular vesicles modulate the cellular status of pericytes. Sci. Rep. 2015, 5, 8505. [Google Scholar] [CrossRef]
  178. Akbar, N.; Digby, J.E.; Cahill, T.J.; Tavare, A.N.; Corbin, A.L.; Saluja, S.; Dawkins, S.; Edgar, L.; Rawlings, N.; Ziberna, K.; et al. Endothelium-derived extracellular vesicles promote splenic monocyte mobilization in myocardial infarction. JCI Insight 2017, 2, e93344. [Google Scholar] [CrossRef] [Green Version]
  179. Arderiu, G.; Pena, E.; Badimon, L. Ischemic tissue released microvesicles induce monocyte reprogramming and increase tissue repair by a tissue factor-dependent mechanism. Cardiovasc. Res. 2021. [Google Scholar] [CrossRef]
  180. Arderiu, G.; Pena, E.; Civit-Urgell, A.; Badimon, L. Endothelium-Released Microvesicles Transport miR-126 That Induces Proangiogenic Reprogramming in Monocytes. Front. Immunol. 2022, 13, 836662. [Google Scholar] [CrossRef]
  181. Arderiu, G.; Pena, E.; Badimon, L. Angiogenic microvascular endothelial cells release microparticles rich in tissue factor that promotes postischemic collateral vessel formation. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 348–357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Leroyer, A.S.; Rautou, P.E.; Silvestre, J.S.; Castier, Y.; Leseche, G.; Devue, C.; Duriez, M.; Brandes, R.P.; Lutgens, E.; Tedgui, A.; et al. CD40 ligand+ microparticles from human atherosclerotic plaques stimulate endothelial proliferation and angiogenesis a potential mechanism for intraplaque neovascularization. J. Am. Coll Cardiol. 2008, 52, 1302–1311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Leroyer, A.S.; Ebrahimian, T.G.; Cochain, C.; Recalde, A.; Blanc-Brude, O.; Mees, B.; Vilar, J.; Tedgui, A.; Levy, B.I.; Chimini, G.; et al. Microparticles from ischemic muscle promotes postnatal vasculogenesis. Circulation 2009, 119, 2808–2817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Lacroix, R.; Sabatier, F.; Mialhe, A.; Basire, A.; Pannell, R.; Borghi, H.; Robert, S.; Lamy, E.; Plawinski, L.; Camoin-Jau, L.; et al. Activation of plasminogen into plasmin at the surface of endothelial microparticles: A mechanism that modulates angiogenic properties of endothelial progenitor cells in vitro. Blood 2007, 110, 2432–2439. [Google Scholar] [CrossRef]
  185. Loyer, X.; Zlatanova, I.; Devue, C.; Yin, M.; Howangyin, K.Y.; Klaihmon, P.; Guerin, C.L.; Kheloufi, M.; Vilar, J.; Zannis, K.; et al. Intra-Cardiac Release of Extracellular Vesicles Shapes Inflammation Following Myocardial Infarction. Circ. Res. 2018, 123, 100–106. [Google Scholar] [CrossRef]
  186. 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] [Green Version]
  187. Bobryshev, Y.V.; Killingsworth, M.C.; Lord, R.S.; Grabs, A.J. Matrix vesicles in the fibrous cap of atherosclerotic plaque: Possible contribution to plaque rupture. J. Cell Mol. Med. 2008, 12, 2073–2082. [Google Scholar] [CrossRef] [Green Version]
  188. Lozito, T.P.; Tuan, R.S. Endothelial cell microparticles act as centers of matrix metalloproteinsase-2 (MMP-2) activation and vascular matrix remodeling. J. Cell Physiol. 2012, 227, 534–549. [Google Scholar] [CrossRef] [Green Version]
  189. Martinez de Lizarrondo, S.; Roncal, C.; Calvayrac, O.; Rodriguez, C.; Varo, N.; Purroy, A.; Lorente, L.; Rodriguez, J.A.; Doeuvre, L.; Hervas-Stubbs, S.; et al. Synergistic effect of thrombin and CD40 ligand on endothelial matrix metalloproteinase-10 expression and microparticle generation in vitro and in vivo. Arterioscler. Thromb. Vasc. Biol. 2012, 32, 1477–1487. [Google Scholar] [CrossRef] [Green Version]
  190. Taraboletti, G.; D’Ascenzo, S.; Borsotti, P.; Giavazzi, R.; Pavan, A.; Dolo, V. Shedding of the matrix metalloproteinases MMP-2, MMP-9, and MT1-MMP as membrane vesicle-associated components by endothelial cells. Am. J. Pathol. 2002, 160, 673–680. [Google Scholar] [CrossRef] [Green Version]
  191. Li, C.J.; Liu, Y.; Chen, Y.; Yu, D.; Williams, K.J.; Liu, M.L. Novel proteolytic microvesicles released from human macrophages after exposure to tobacco smoke. Am. J. Pathol. 2013, 182, 1552–1562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Folkesson, M.; Li, C.; Frebelius, S.; Swedenborg, J.; Wagsater, D.; Williams, K.J.; Eriksson, P.; Roy, J.; Liu, M.L. Proteolytically active ADAM10 and ADAM17 carried on membrane microvesicles in human abdominal aortic aneurysms. Thromb. Haemost. 2015, 114, 1165–1174. [Google Scholar] [CrossRef]
  193. Gasser, O.; Hess, C.; Miot, S.; Deon, C.; Sanchez, J.C.; Schifferli, J.A. Characterisation and properties of ectosomes released by human polymorphonuclear neutrophils. Exp. Cell Res. 2003, 285, 243–257. [Google Scholar] [CrossRef]
  194. Falati, S.; Gross, P.; Merrill-Skoloff, G.; Furie, B.C.; Furie, B. Real-time in vivo imaging of platelets, tissue factor and fibrin during arterial thrombus formation in the mouse. Nat. Med. 2002, 8, 1175–1181. [Google Scholar] [CrossRef] [PubMed]
  195. Oggero, S.; de Gaetano, M.; Marcone, S.; Fitzsimons, S.; Pinto, A.L.; Ikramova, D.; Barry, M.; Burke, D.; Montero-Melendez, T.; Cooper, D.; et al. Extracellular vesicles from monocyte/platelet aggregates modulate human atherosclerotic plaque reactivity. J. Extracell Vesicles 2021, 10, 12084. [Google Scholar] [CrossRef]
  196. Bohm, J.K.; Schafer, N.; Maegele, M.; Stumpges, B.; Bauerfeind, U.; Caspers, M. Plasmatic and cell-based enhancement by microparticles originated from platelets and endothelial cells under simulated in vitro conditions of a dilutional coagulopathy. Scand. J. Trauma Resusc. Emerg. Med. 2021, 29, 38. [Google Scholar] [CrossRef]
  197. Aharon, A.; Tamari, T.; Brenner, B. Monocyte-derived microparticles and exosomes induce procoagulant and apoptotic effects on endothelial cells. Thromb. Haemost. 2008, 100, 878–885. [Google Scholar] [CrossRef]
  198. Lacroix, R.; Plawinski, L.; Robert, S.; Doeuvre, L.; Sabatier, F.; Martinez de Lizarrondo, S.; Mezzapesa, A.; Anfosso, F.; Leroyer, A.S.; Poullin, P.; et al. Leukocyte- and endothelial-derived microparticles: A circulating source for fibrinolysis. Haematologica 2012, 97, 1864–1872. [Google Scholar] [CrossRef]
  199. Steppich, B.; Mattisek, C.; Sobczyk, D.; Kastrati, A.; Schomig, A.; Ott, I. Tissue factor pathway inhibitor on circulating microparticles in acute myocardial infarction. Thromb. Haemost. 2005, 93, 35–39. [Google Scholar] [CrossRef]
  200. Biro, E.; Sturk-Maquelin, K.N.; Vogel, G.M.; Meuleman, D.G.; Smit, M.J.; Hack, C.E.; Sturk, A.; Nieuwland, R. Human cell-derived microparticles promote thrombus formation in vivo in a tissue factor-dependent manner. J. Thromb. Haemost. 2003, 1, 2561–2568. [Google Scholar] [CrossRef] [Green Version]
  201. Van Der Meijden, P.E.; Van Schilfgaarde, M.; Van Oerle, R.; Renne, T.; ten Cate, H.; Spronk, H.M. Platelet- and erythrocyte-derived microparticles trigger thrombin generation via factor XIIa. J. Thromb. Haemost. 2012, 10, 1355–1362. [Google Scholar] [CrossRef] [PubMed]
  202. Chou, J.; Mackman, N.; Merrill-Skoloff, G.; Pedersen, B.; Furie, B.C.; Furie, B. Hematopoietic cell-derived microparticle tissue factor contributes to fibrin formation during thrombus propagation. Blood 2004, 104, 3190–3197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Hugel, B.; Socie, G.; Vu, T.; Toti, F.; Gluckman, E.; Freyssinet, J.M.; Scrobohaci, M.L. Elevated levels of circulating procoagulant microparticles in patients with paroxysmal nocturnal hemoglobinuria and aplastic anemia. Blood 1999, 93, 3451–3456. [Google Scholar] [CrossRef] [PubMed]
  204. Van Beers, E.J.; Schaap, M.C.; Berckmans, R.J.; Nieuwland, R.; Sturk, A.; van Doormaal, F.F.; Meijers, J.C.; Biemond, B.J.; group, C.s. Circulating erythrocyte-derived microparticles are associated with coagulation activation in sickle cell disease. Haematologica 2009, 94, 1513–1519. [Google Scholar] [CrossRef] [Green Version]
  205. Suades, R.; Padro, T.; Vilahur, G.; Badimon, L. Circulating and platelet-derived microparticles in human blood enhance thrombosis on atherosclerotic plaques. Thromb. Haemost. 2012, 108, 1208–1219. [Google Scholar] [CrossRef]
  206. Tripisciano, C.; Weiss, R.; Eichhorn, T.; Spittler, A.; Heuser, T.; Fischer, M.B.; Weber, V. Different Potential of Extracellular Vesicles to Support Thrombin Generation: Contributions of Phosphatidylserine, Tissue Factor, and Cellular Origin. Sci. Rep. 2017, 7, 6522. [Google Scholar] [CrossRef]
  207. Suades, R.; Padro, T.; Vilahur, G.; Martin-Yuste, V.; Sabate, M.; Sans-Rosello, J.; Sionis, A.; Badimon, L. Growing thrombi release increased levels of CD235a(+) microparticles and decreased levels of activated platelet-derived microparticles. Validation in ST-elevation myocardial infarction patients. J. Thromb. Haemost. 2015, 13, 1776–1786. [Google Scholar] [CrossRef]
  208. Muller, I.; Klocke, A.; Alex, M.; Kotzsch, M.; Luther, T.; Morgenstern, E.; Zieseniss, S.; Zahler, S.; Preissner, K.; Engelmann, B. Intravascular tissue factor initiates coagulation via circulating microvesicles and platelets. FASEB J. 2003, 17, 476–478. [Google Scholar] [CrossRef] [Green Version]
  209. Geddings, J.E.; Hisada, Y.; Boulaftali, Y.; Getz, T.M.; Whelihan, M.; Fuentes, R.; Dee, R.; Cooley, B.C.; Key, N.S.; Wolberg, A.S.; et al. Tissue factor-positive tumor microvesicles activate platelets and enhance thrombosis in mice. J. Thromb. Haemost. 2016, 14, 153–166. [Google Scholar] [CrossRef] [Green Version]
  210. Zwicker, J.I.; Trenor, C.C., 3rd; Furie, B.C.; Furie, B. Tissue factor-bearing microparticles and thrombus formation. Arterioscler. Thromb. Vasc. Biol. 2011, 31, 728–733. [Google Scholar] [CrossRef] [Green Version]
  211. Scholz, T.; Temmler, U.; Krause, S.; Heptinstall, S.; Losche, W. Transfer of tissue factor from platelets to monocytes: Role of platelet-derived microvesicles and CD62P. Thromb. Haemost. 2002, 88, 1033–1038. [Google Scholar] [CrossRef] [PubMed]
  212. Collier, M.E.W.; Ettelaie, C.; Goult, B.T.; Maraveyas, A.; Goodall, A.H. Investigation of the Filamin A-Dependent Mechanisms of Tissue Factor Incorporation into Microvesicles. Thromb. Haemost. 2017, 117, 2034–2044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Stojkovic, S.; Thulin, A.; Hell, L.; Thaler, B.; Rauscher, S.; Baumgartner, J.; Groger, M.; Ay, C.; Demyanets, S.; Neumayer, C.; et al. IL-33 stimulates the release of procoagulant microvesicles from human monocytes and differentially increases tissue factor in human monocyte subsets. Thromb. Haemost. 2017, 117, 1379–1390. [Google Scholar] [CrossRef] [PubMed]
  214. Sabatier, F.; Roux, V.; Anfosso, F.; Camoin, L.; Sampol, J.; Dignat-George, F. Interaction of endothelial microparticles with monocytic cells in vitro induces tissue factor-dependent procoagulant activity. Blood 2002, 99, 3962–3970. [Google Scholar] [CrossRef]
  215. Del Conde, I.; Shrimpton, C.N.; Thiagarajan, P.; Lopez, J.A. Tissue-factor-bearing microvesicles arise from lipid rafts and fuse with activated platelets to initiate coagulation. Blood 2005, 106, 1604–1611. [Google Scholar] [CrossRef]
  216. Lopez-Vilchez, I.; Escolar, G.; Diaz-Ricart, M.; Fuste, B.; Galan, A.M.; White, J.G. Tissue factor-enriched vesicles are taken up by platelets and induce platelet aggregation in the presence of factor VIIa. Thromb. Haemost. 2007, 97, 202–211. [Google Scholar] [CrossRef]
  217. Gross, P.L.; Furie, B.C.; Merrill-Skoloff, G.; Chou, J.; Furie, B. Leukocyte-versus microparticle-mediated tissue factor transfer during arteriolar thrombus development. J. Leukoc. Biol. 2005, 78, 1318–1326. [Google Scholar] [CrossRef]
  218. Sinauridze, E.I.; Kireev, D.A.; Popenko, N.Y.; Pichugin, A.V.; Panteleev, M.A.; Krymskaya, O.V.; Ataullakhanov, F.I. Platelet microparticle membranes have 50- to 100-fold higher specific procoagulant activity than activated platelets. Thromb. Haemost. 2007, 97, 425–434. [Google Scholar]
  219. Berckmans, R.J.; Nieuwland, R.; Boing, A.N.; Romijn, F.P.; Hack, C.E.; Sturk, A. Cell-derived microparticles circulate in healthy humans and support low grade thrombin generation. Thromb. Haemost. 2001, 85, 639–646. [Google Scholar]
  220. Zhao, L.; Bi, Y.; Kou, J.; Shi, J.; Piao, D. Phosphatidylserine exposing-platelets and microparticles promote procoagulant activity in colon cancer patients. J. Exp. Clin. Cancer Res. 2016, 35, 54. [Google Scholar] [CrossRef] [Green Version]
  221. Falati, S.; Liu, Q.; Gross, P.; Merrill-Skoloff, G.; Chou, J.; Vandendries, E.; Celi, A.; Croce, K.; Furie, B.C.; Furie, B. Accumulation of tissue factor into developing thrombi in vivo is dependent upon microparticle P-selectin glycoprotein ligand 1 and platelet P-selectin. J. Exp. Med. 2003, 197, 1585–1598. [Google Scholar] [CrossRef] [PubMed]
  222. Keuren, J.F.; Magdeleyns, E.J.; Bennaghmouch, A.; Bevers, E.M.; Curvers, J.; Lindhout, T. Microparticles adhere to collagen type I, fibrinogen, von Willebrand factor and surface immobilised platelets at physiological shear rates. Br. J. Haematol. 2007, 138, 527–533. [Google Scholar] [CrossRef]
  223. Siljander, P.; Carpen, O.; Lassila, R. Platelet-derived microparticles associate with fibrin during thrombosis. Blood 1996, 87, 4651–4663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Merten, M.; Pakala, R.; Thiagarajan, P.; Benedict, C.R. Platelet microparticles promote platelet interaction with subendothelial matrix in a glycoprotein IIb/IIIa-dependent mechanism. Circulation 1999, 99, 2577–2582. [Google Scholar] [CrossRef] [Green Version]
  225. Raturi, A.; Miersch, S.; Hudson, J.W.; Mutus, B. Platelet microparticle-associated protein disulfide isomerase promotes platelet aggregation and inactivates insulin. Biochim. Biophys. Acta 2008, 1778, 2790–2796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Sims, P.J.; Faioni, E.M.; Wiedmer, T.; Shattil, S.J. Complement proteins C5b-9 cause release of membrane vesicles from the platelet surface that are enriched in the membrane receptor for coagulation factor Va and express prothrombinase activity. J. Biol. Chem. 1988, 263, 18205–18212. [Google Scholar] [CrossRef]
  227. Nomura, S.; Komiyama, Y.; Murakami, T.; Funatsu, A.; Kokawa, T.; Sugo, T.; Matsuda, M.; Yasunaga, K. Flow cytometric analysis of surface membrane proteins on activated platelets and platelet-derived microparticles from healthy and thrombasthenic individuals. Int. J. Hematol. 1993, 58, 203–212. [Google Scholar]
  228. Barry, O.P.; Pratico, D.; Lawson, J.A.; FitzGerald, G.A. Transcellular activation of platelets and endothelial cells by bioactive lipids in platelet microparticles. J. Clin. Investig. 1997, 99, 2118–2127. [Google Scholar] [CrossRef] [Green Version]
  229. Qu, M.; Zou, X.; Fang, F.; Wang, S.; Xu, L.; Zeng, Q.; Fan, Z.; Chen, L.; Yue, W.; Xie, X.; et al. Platelet-derived microparticles enhance megakaryocyte differentiation and platelet generation via miR-1915-3p. Nat. Commun. 2020, 11, 4964. [Google Scholar] [CrossRef]
  230. Suades, R.; Padro, T.; Vilahur, G.; Badimon, L. Platelet-released extracellular vesicles: The effects of thrombin activation. Cell Mol. Life Sci. 2022, 79, 190. [Google Scholar] [CrossRef]
  231. Banfi, C.; Brioschi, M.; Wait, R.; Begum, S.; Gianazza, E.; Pirillo, A.; Mussoni, L.; Tremoli, E. Proteome of endothelial cell-derived procoagulant microparticles. Proteomics 2005, 5, 4443–4455. [Google Scholar] [CrossRef] [PubMed]
  232. Landers-Ramos, R.Q.; Addison, O.A.; Beamer, B.; Katzel, L.I.; Blumenthal, J.B.; Robinson, S.; Hagberg, J.M.; Prior, S.J. Circulating microparticle concentrations across acute and chronic cardiovascular disease conditions. Physiol. Rep. 2020, 8, e14534. [Google Scholar] [CrossRef]
  233. Marei, I.; Chidiac, O.; Thomas, B.; Pasquier, J.; Dargham, S.; Robay, A.; Vakayil, M.; Jameesh, M.; Triggle, C.; Rafii, A.; et al. Angiogenic content of microparticles in patients with diabetes and coronary artery disease predicts networks of endothelial dysfunction. Cardiovasc. Diabetol. 2022, 21, 17. [Google Scholar] [CrossRef] [PubMed]
  234. Wang, Y.; Chen, L.M.; Liu, M.L. Microvesicles and diabetic complications—Novel mediators, potential biomarkers and therapeutic targets. Acta Pharmacol. Sin. 2014, 35, 433–443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. La Salvia, S.; Musante, L.; Lannigan, J.; Gigliotti, J.C.; Le, T.H.; Erdbrugger, U. T cell-derived extracellular vesicles are elevated in essential HTN. Am. J. Physiol. Renal Physiol. 2020, 319, F868–F875. [Google Scholar] [CrossRef] [PubMed]
  236. Alexandru, N.; Badila, E.; Weiss, E.; Cochior, D.; Stepien, E.; Georgescu, A. Vascular complications in diabetes: Microparticles and microparticle associated microRNAs as active players. Biochem. Biophys. Res. Commun. 2016, 472, 1–10. [Google Scholar] [CrossRef] [PubMed]
  237. Jung, K.H.; Chu, K.; Lee, S.T.; Park, H.K.; Bahn, J.J.; Kim, D.H.; Kim, J.H.; Kim, M.; Kun Lee, S.; Roh, J.K. Circulating endothelial microparticles as a marker of cerebrovascular disease. Ann. Neurol. 2009, 66, 191–199. [Google Scholar] [CrossRef]
  238. Bernal-Mizrachi, L.; Jy, W.; Jimenez, J.J.; Pastor, J.; Mauro, L.M.; Horstman, L.L.; de Marchena, E.; Ahn, Y.S. High levels of circulating endothelial microparticles in patients with acute coronary syndromes. Am. Heart J. 2003, 145, 962–970. [Google Scholar] [CrossRef]
  239. Montoro-Garcia, S.; Shantsila, E.; Tapp, L.D.; Lopez-Cuenca, A.; Romero, A.I.; Hernandez-Romero, D.; Orenes-Pinero, E.; Manzano-Fernandez, S.; Valdes, M.; Marin, F.; et al. Small-size circulating microparticles in acute coronary syndromes: Relevance to fibrinolytic status, reparative markers and outcomes. Atherosclerosis 2013, 227, 313–322. [Google Scholar] [CrossRef]
  240. Nozaki, T.; Sugiyama, S.; Koga, H.; Sugamura, K.; Ohba, K.; Matsuzawa, Y.; Sumida, H.; Matsui, K.; Jinnouchi, H.; Ogawa, H. Significance of a multiple biomarkers strategy including endothelial dysfunction to improve risk stratification for cardiovascular events in patients at high risk for coronary heart disease. J. Am. Coll Cardiol. 2009, 54, 601–608. [Google Scholar] [CrossRef] [Green Version]
  241. Jansen, F.; Yang, X.; Proebsting, S.; Hoelscher, M.; Przybilla, D.; Baumann, K.; Schmitz, T.; Dolf, A.; Endl, E.; Franklin, B.S.; et al. MicroRNA expression in circulating microvesicles predicts cardiovascular events in patients with coronary artery disease. J. Am. Heart Assoc. 2014, 3, e001249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Suades, R.; Padro, T.; Crespo, J.; Sionis, A.; Alonso, R.; Mata, P.; Badimon, L. Liquid Biopsy of Extracellular Microvesicles Predicts Future Major Ischemic Events in Genetically Characterized Familial Hypercholesterolemia Patients. Arterioscler. Thromb. Vasc. Biol. 2019, 39, 1172–1181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Huo, S.; Krankel, N.; Nave, A.H.; Sperber, P.S.; Rohmann, J.L.; Piper, S.K.; Heuschmann, P.U.; Landmesser, U.; Endres, M.; Siegerink, B.; et al. Endothelial and Leukocyte-Derived Microvesicles and Cardiovascular Risk After Stroke: PROSCIS-B. Neurology 2021, 96, e937–e946. [Google Scholar] [CrossRef] [PubMed]
  244. Chiva-Blanch, G.; Bratseth, V.; Ritschel, V.; Andersen, G.O.; Halvorsen, S.; Eritsland, J.; Arnesen, H.; Badimon, L.; Seljeflot, I. Monocyte-derived circulating microparticles (CD14(+), CD14(+)/CD11b(+) and CD14(+)/CD142(+)) are related to long-term prognosis for cardiovascular mortality in STEMI patients. Int. J. Cardiol. 2017, 227, 876–881. [Google Scholar] [CrossRef] [Green Version]
  245. Suades, R.; Padro, T.; Crespo, J.; Ramaiola, I.; Martin-Yuste, V.; Sabate, M.; Sans-Rosello, J.; Sionis, A.; Badimon, L. Circulating microparticle signature in coronary and peripheral blood of ST elevation myocardial infarction patients in relation to pain-to-PCI elapsed time. Int. J. Cardiol. 2016, 202, 378–387. [Google Scholar] [CrossRef]
  246. Sarlon-Bartoli, G.; Bennis, Y.; Lacroix, R.; Piercecchi-Marti, M.D.; Bartoli, M.A.; Arnaud, L.; Mancini, J.; Boudes, A.; Sarlon, E.; Thevenin, B.; et al. Plasmatic level of leukocyte-derived microparticles is associated with unstable plaque in asymptomatic patients with high-grade carotid stenosis. J. Am. Coll Cardiol. 2013, 62, 1436–1441. [Google Scholar] [CrossRef] [Green Version]
  247. Morel, O.; Hugel, B.; Jesel, L.; Lanza, F.; Douchet, M.P.; Zupan, M.; Chauvin, M.; Cazenave, J.P.; Freyssinet, J.M.; Toti, F. Sustained elevated amounts of circulating procoagulant membrane microparticles and soluble GPV after acute myocardial infarction in diabetes mellitus. Thromb. Haemost. 2004, 91, 345–353. [Google Scholar] [CrossRef]
  248. Mallat, Z.; Benamer, H.; Hugel, B.; Benessiano, J.; Steg, P.G.; Freyssinet, J.M.; Tedgui, A. Elevated levels of shed membrane microparticles with procoagulant potential in the peripheral circulating blood of patients with acute coronary syndromes. Circulation 2000, 101, 841–843. [Google Scholar] [CrossRef] [Green Version]
  249. Van der Zee, P.M.; Biro, E.; Ko, Y.; de Winter, R.J.; Hack, C.E.; Sturk, A.; Nieuwland, R. P-selectin- and CD63-exposing platelet microparticles reflect platelet activation in peripheral arterial disease and myocardial infarction. Clin. Chem. 2006, 52, 657–664. [Google Scholar] [CrossRef]
  250. Stepien, E.; Stankiewicz, E.; Zalewski, J.; Godlewski, J.; Zmudka, K.; Wybranska, I. Number of microparticles generated during acute myocardial infarction and stable angina correlates with platelet activation. Arch. Med. Res. 2012, 43, 31–35. [Google Scholar] [CrossRef]
  251. Katopodis, J.N.; Kolodny, L.; Jy, W.; Horstman, L.L.; De Marchena, E.J.; Tao, J.G.; Haynes, D.H.; Ahn, Y.S. Platelet microparticles and calcium homeostasis in acute coronary ischemias. Am. J. Hematol. 1997, 54, 95–101. [Google Scholar] [CrossRef]
  252. Giannopoulos, G.; Oudatzis, G.; Paterakis, G.; Synetos, A.; Tampaki, E.; Bouras, G.; Hahalis, G.; Alexopoulos, D.; Tousoulis, D.; Cleman, M.W.; et al. Red blood cell and platelet microparticles in myocardial infarction patients treated with primary angioplasty. Int. J. Cardiol. 2014, 176, 145–150. [Google Scholar] [CrossRef] [PubMed]
  253. Biasucci, L.M.; Porto, I.; Di Vito, L.; De Maria, G.L.; Leone, A.M.; Tinelli, G.; Tritarelli, A.; Di Rocco, G.; Snider, F.; Capogrossi, M.C.; et al. Differences in microparticle release in patients with acute coronary syndrome and stable angina. Circ. J. 2012, 76, 2174–2182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Min, P.K.; Kim, J.Y.; Chung, K.H.; Lee, B.K.; Cho, M.; Lee, D.L.; Hong, S.Y.; Choi, E.Y.; Yoon, Y.W.; Hong, B.K.; et al. Local increase in microparticles from the aspirate of culprit coronary arteries in patients with ST-segment elevation myocardial infarction. Atherosclerosis 2013, 227, 323–328. [Google Scholar] [CrossRef] [PubMed]
  255. Porto, I.; Biasucci, L.M.; De Maria, G.L.; Leone, A.M.; Niccoli, G.; Burzotta, F.; Trani, C.; Tritarelli, A.; Vergallo, R.; Liuzzo, G.; et al. Intracoronary microparticles and microvascular obstruction in patients with ST elevation myocardial infarction undergoing primary percutaneous intervention. Eur. Heart J. 2012, 33, 2928–2938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Werner, N.; Wassmann, S.; Ahlers, P.; Kosiol, S.; Nickenig, G. Circulating CD31+/annexin V+ apoptotic microparticles correlate with coronary endothelial function in patients with coronary artery disease. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 112–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Van Ierssel, S.H.; Hoymans, V.Y.; Van Craenenbroeck, E.M.; Van Tendeloo, V.F.; Vrints, C.J.; Jorens, P.G.; Conraads, V.M. Endothelial microparticles (EMP) for the assessment of endothelial function: An in vitro and in vivo study on possible interference of plasma lipids. PLoS ONE 2012, 7, e31496. [Google Scholar] [CrossRef] [Green Version]
  258. Sinning, J.M.; Losch, J.; Walenta, K.; Bohm, M.; Nickenig, G.; Werner, N. Circulating CD31+/Annexin V+ microparticles correlate with cardiovascular outcomes. Eur. Heart J. 2011, 32, 2034–2041. [Google Scholar] [CrossRef] [Green Version]
  259. Lee, S.T.; Chu, K.; Jung, K.H.; Kim, J.M.; Moon, H.J.; Bahn, J.J.; Im, W.S.; Sunwoo, J.; Moon, J.; Kim, M.; et al. Circulating CD62E+ microparticles and cardiovascular outcomes. PLoS ONE 2012, 7, e35713. [Google Scholar] [CrossRef]
  260. Amabile, N.; Heiss, C.; Real, W.M.; Minasi, P.; McGlothlin, D.; Rame, E.J.; Grossman, W.; De Marco, T.; Yeghiazarians, Y. Circulating endothelial microparticle levels predict hemodynamic severity of pulmonary hypertension. Am. J. Respir. Crit. Care Med. 2008, 177, 1268–1275. [Google Scholar] [CrossRef]
  261. Zacharia, E.; Antonopoulos, A.S.; Oikonomou, E.; Papageorgiou, N.; Pallantza, Z.; Miliou, A.; Mystakidi, V.C.; Simantiris, S.; Kriebardis, A.; Orologas, N.; et al. Plasma signature of apoptotic microvesicles is associated with endothelial dysfunction and plaque rupture in acute coronary syndromes. J. Mol. Cell Cardiol. 2020, 138, 110–114. [Google Scholar] [CrossRef] [PubMed]
  262. Bernal-Mizrachi, L.; Jy, W.; Fierro, C.; Macdonough, R.; Velazques, H.A.; Purow, J.; Jimenez, J.J.; Horstman, L.L.; Ferreira, A.; de Marchena, E.; et al. Endothelial microparticles correlate with high-risk angiographic lesions in acute coronary syndromes. Int. J. Cardiol. 2004, 97, 439–446. [Google Scholar] [CrossRef] [PubMed]
  263. Chiva-Blanch, G.; Padro, T.; Alonso, R.; Crespo, J.; Perez de Isla, L.; Mata, P.; Badimon, L. Liquid Biopsy of Extracellular Microvesicles Maps Coronary Calcification and Atherosclerotic Plaque in Asymptomatic Patients With Familial Hypercholesterolemia. Arterioscler. Thromb. Vasc. Biol. 2019, 39, 945–955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Ueba, T.; Nomura, S.; Inami, N.; Nishikawa, T.; Kajiwara, M.; Iwata, R.; Yamashita, K. Plasma level of platelet-derived microparticles is associated with coronary heart disease risk score in healthy men. J. Atheroscler. Thromb. 2010, 17, 342–349. [Google Scholar] [CrossRef] [Green Version]
  265. Chiva-Blanch, G.; Suades, R.; Crespo, J.; Vilahur, G.; Arderiu, G.; Padro, T.; Corella, D.; Salas-Salvado, J.; Aros, F.; Martinez-Gonzalez, M.A.; et al. CD3(+)/CD45(+) and SMA-alpha(+) circulating microparticles are increased in individuals at high cardiovascular risk who will develop a major cardiovascular event. Int. J. Cardiol. 2016, 208, 147–149. [Google Scholar] [CrossRef] [Green Version]
  266. Suades, R.; Padro, T.; Alonso, R.; Mata, P.; Badimon, L. High levels of TSP1+/CD142+ platelet-derived microparticles characterise young patients with high cardiovascular risk and subclinical atherosclerosis. Thromb. Haemost. 2015, 114, 1310–1321. [Google Scholar] [CrossRef]
  267. Lazaridis, A.; Gavriilaki, E.; Nikolaidou, B.; Yiannaki, E.; Dolgyras, P.; Anyfanti, P.; Triantafyllou, A.; Koletsos, N.; Tzimos, C.; Markala, D.; et al. A study of endothelial and platelet microvesicles across different hypertension phenotypes. J. Hum. Hypertens 2021. [Google Scholar] [CrossRef]
  268. Csongradi, E.; Nagy, B., Jr.; Fulop, T.; Varga, Z.; Karanyi, Z.; Magyar, M.T.; Olah, L.; Papp, M.; Facsko, A.; Kappelmayer, J.; et al. Increased levels of platelet activation markers are positively associated with carotid wall thickness and other atherosclerotic risk factors in obese patients. Thromb. Haemost. 2011, 106, 683–692. [Google Scholar] [CrossRef] [Green Version]
  269. Mobarrez, F.; Antoniewicz, L.; Bosson, J.A.; Kuhl, J.; Pisetsky, D.S.; Lundback, M. The effects of smoking on levels of endothelial progenitor cells and microparticles in the blood of healthy volunteers. PLoS ONE 2014, 9, e90314. [Google Scholar] [CrossRef] [Green Version]
  270. Wekesa, A.L.; Cross, K.S.; O’Donovan, O.; Dowdall, J.F.; O’Brien, O.; Doyle, M.; Byrne, L.; Phelan, J.P.; Ross, M.D.; Landers, R.; et al. Predicting carotid artery disease and plaque instability from cell-derived microparticles. Eur. J. Vasc. Endovasc. Surg. 2014, 48, 489–495. [Google Scholar] [CrossRef] [Green Version]
  271. Sansone, R.; Baaken, M.; Horn, P.; Schuler, D.; Westenfeld, R.; Amabile, N.; Kelm, M.; Heiss, C. Endothelial microparticles and vascular parameters in subjects with and without arterial hypertension and coronary artery disease. Data Brief. 2018, 19, 495–500. [Google Scholar] [CrossRef] [PubMed]
  272. Wang, B.; Li, T.; Han, X.; Li, Y.; Cheng, W.; Wang, L.; Lu, Z.; Yang, J.; Zhao, M. The Level of Circulating Microparticles in Patients with Coronary Heart Disease: A Systematic Review and Meta-Analysis. J. Cardiovasc. Transl. Res. 2020, 13, 702–712. [Google Scholar] [CrossRef] [PubMed]
  273. Cui, Y.; Zheng, L.; Jiang, M.; Jia, R.; Zhang, X.; Quan, Q.; Du, G.; Shen, D.; Zhao, X.; Sun, W.; et al. Circulating microparticles in patients with coronary heart disease and its correlation with interleukin-6 and C-reactive protein. Mol. Biol. Rep. 2013, 40, 6437–6442. [Google Scholar] [CrossRef] [PubMed]
  274. Cheng, G.; Shan, X.F.; Wang, X.L.; Dong, W.W.; Li, Z.; Liu, X.H.; Zhang, W.; Xing, K.; Chang, F.J. Endothelial damage effects of circulating microparticles from patients with stable angina are reduced by aspirin through ERK/p38 MAPKs pathways. Cardiovasc.Ther. 2017, 35, e12273. [Google Scholar] [CrossRef]
  275. Vagida, M.S.; Arakelyan, A.; Lebedeva, A.M.; Grivel, J.C.; Shpektor, A.V.; Vasilieva, E.Y.; Margolis, L.B. Analysis of Extracellular Vesicles Using Magnetic Nanoparticles in Blood of Patients with Acute Coronary Syndrome. Biochemistry 2016, 81, 382–391. [Google Scholar] [CrossRef] [Green Version]
  276. Fan, Y.; Wang, L.; Li, Y.; Yin, Z.; Xu, Z.; Wang, C. Quantification of endothelial microparticles on modified cytometric bead assay and prognosis in chest pain patients. Circ. J. 2014, 78, 206–214. [Google Scholar] [CrossRef] [Green Version]
  277. Empana, J.P.; Boulanger, C.M.; Tafflet, M.; Renard, J.M.; Leroyer, A.S.; Varenne, O.; Prugger, C.; Silvain, J.; Tedgui, A.; Cariou, A.; et al. Microparticles and sudden cardiac death due to coronary occlusion. The TIDE (Thrombus and Inflammation in sudden DEath) study. Eur. Heart J. Acute Cardiovasc. Care 2015, 4, 28–36. [Google Scholar] [CrossRef]
  278. Mortberg, J.; Lundwall, K.; Mobarrez, F.; Wallen, H.; Jacobson, S.H.; Spaak, J. Increased concentrations of platelet- and endothelial-derived microparticles in patients with myocardial infarction and reduced renal function—A descriptive study. BMC Nephrol. 2019, 20, 71. [Google Scholar] [CrossRef]
  279. Gasecka, A.; Nieuwland, R.; Budnik, M.; Dignat-George, F.; Eyileten, C.; Harrison, P.; Lacroix, R.; Leroyer, A.; Opolski, G.; Pluta, K.; et al. Ticagrelor attenuates the increase of extracellular vesicle concentrations in plasma after acute myocardial infarction compared to clopidogrel. J. Thromb. Haemost. 2020, 18, 609–623. [Google Scholar] [CrossRef] [Green Version]
  280. Christersson, C.; Thulin, A.; Siegbahn, A. Microparticles during long-term follow-up after acute myocardial infarction. Association to atherosclerotic burden and risk of cardiovascular events. Thromb. Haemost. 2017, 117, 1571–1581. [Google Scholar] [CrossRef]
  281. Li, P.; Qin, C. Elevated circulating VE-cadherin+CD144+endothelial microparticles in ischemic cerebrovascular disease. Thromb. Res. 2015, 135, 375–381. [Google Scholar] [CrossRef] [PubMed]
  282. Chiva-Blanch, G.; Suades, R.; Crespo, J.; Pena, E.; Padro, T.; Jimenez-Xarrie, E.; Marti-Fabregas, J.; Badimon, L. Microparticle Shedding from Neural Progenitor Cells and Vascular Compartment Cells Is Increased in Ischemic Stroke. PLoS ONE 2016, 11, e0148176. [Google Scholar] [CrossRef] [PubMed]
  283. Bivard, A.; Lincz, L.F.; Maquire, J.; Parsons, M.; Levi, C. Platelet microparticles: A biomarker for recanalization in rtPA-treated ischemic stroke patients. Ann. Clin. Transl. Neurol. 2017, 4, 175–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Saenz-Pipaon, G.; San Martin, P.; Planell, N.; Maillo, A.; Ravassa, S.; Vilas-Zornoza, A.; Martinez-Aguilar, E.; Rodriguez, J.A.; Alameda, D.; Lara-Astiaso, D.; et al. Functional and transcriptomic analysis of extracellular vesicles identifies calprotectin as a new prognostic marker in peripheral arterial disease (PAD). J. Extracell Vesicles 2020, 9, 1729646. [Google Scholar] [CrossRef]
  285. Giarretta, I.; Gatto, I.; Marcantoni, M.; Lupi, G.; Tonello, D.; Gaetani, E.; Pitocco, D.; Iezzi, R.; Truma, A.; Porfidia, A.; et al. Microparticles Carrying Sonic Hedgehog Are Increased in Humans with Peripheral Artery Disease. Int. J. Mol. Sci 2018, 19, 3954. [Google Scholar] [CrossRef] [Green Version]
  286. Crawford, J.R.; Trial, J.; Nambi, V.; Hoogeveen, R.C.; Taffet, G.E.; Entman, M.L. Plasma Levels of Endothelial Microparticles Bearing Monomeric C-reactive Protein are Increased in Peripheral Artery Disease. J. Cardiovasc. Transl. Res. 2016, 9, 184–193. [Google Scholar] [CrossRef] [Green Version]
  287. Jamaly, S.; Basavaraj, M.G.; Starikova, I.; Olsen, R.; Braekkan, S.K.; Hansen, J.B. Elevated plasma levels of P-selectin glycoprotein ligand-1-positive microvesicles in patients with unprovoked venous thromboembolism. J. Thromb. Haemost. 2018. [Google Scholar] [CrossRef]
  288. Thulin, A.; Lindback, J.; Granger, C.B.; Wallentin, L.; Lind, L.; Siegbahn, A. Extracellular vesicles in atrial fibrillation and stroke. Thromb. Res. 2020, 193, 180–189. [Google Scholar] [CrossRef]
  289. Anyfanti, P.; Gavriilaki, E.; Nikolaidou, B.; Yiannaki, E.; Lazaridis, A.; Papadopoulos, N.; Douma, S.; Doumas, M.; Gkaliagkousi, E. Patients with autoimmune chronic inflammatory diseases present increased biomarkers of thromboinflammation and endothelial dysfunction in the absence of flares and cardiovascular comorbidities. J. Thromb. Thrombolysis 2022, 53, 10–16. [Google Scholar] [CrossRef]
  290. Gustafson, C.M.; Shepherd, A.J.; Miller, V.M.; Jayachandran, M. Age- and sex-specific differences in blood-borne microvesicles from apparently healthy humans. Biol. Sex. Differ. 2015, 6, 10. [Google Scholar] [CrossRef] [Green Version]
  291. Baek, R.; Varming, K.; Jorgensen, M.M. Does smoking, age or gender affect the protein phenotype of extracellular vesicles in plasma? Transfus Apher. Sci. 2016, 55, 44–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Condrat, C.E.; Varlas, V.N.; Duica, F.; Antoniadis, P.; Danila, C.A.; Cretoiu, D.; Suciu, N.; Cretoiu, S.M.; Voinea, S.C. Pregnancy-Related Extracellular Vesicles Revisited. Int. J. Mol. Sci 2021, 22, 3904. [Google Scholar] [CrossRef] [PubMed]
  293. Aboulgheit, A.; Potz, B.A.; Scrimgeour, L.A.; Karbasiafshar, C.; Shi, G.; Zhang, Z.; Machan, J.T.; Schorl, C.; Brodsky, A.S.; Braga, K.; et al. Effects of High Fat Versus Normal Diet on Extracellular Vesicle-Induced Angiogenesis in a Swine Model of Chronic Myocardial Ischemia. J. Am. Heart Assoc. 2021, 10, e017437. [Google Scholar] [CrossRef] [PubMed]
  294. Brahmer, A.; Neuberger, E.; Esch-Heisser, L.; Haller, N.; Jorgensen, M.M.; Baek, R.; Mobius, W.; Simon, P.; Kramer-Albers, E.M. Platelets, endothelial cells and leukocytes contribute to the exercise-triggered release of extracellular vesicles into the circulation. J. Extracell Vesicles 2019, 8, 1615820. [Google Scholar] [CrossRef] [PubMed]
  295. Rigamonti, A.E.; Bollati, V.; Pergoli, L.; Iodice, S.; De Col, A.; Tamini, S.; Cicolini, S.; Tringali, G.; De Micheli, R.; Cella, S.G.; et al. Effects of an acute bout of exercise on circulating extracellular vesicles: Tissue-, sex-, and BMI-related differences. Int. J. Obes. 2020, 44, 1108–1118. [Google Scholar] [CrossRef]
  296. Vechetti, I.J., Jr.; Valentino, T.; Mobley, C.B.; McCarthy, J.J. The role of extracellular vesicles in skeletal muscle and systematic adaptation to exercise. J. Physiol. 2021, 599, 845–861. [Google Scholar] [CrossRef]
  297. Estrada, A.L.; Valenti, Z.J.; Hehn, G.; Amorese, A.J.; Williams, N.S.; Balestrieri, N.P.; Deighan, C.; Allen, C.P.; Spangenburg, E.E.; Kruh-Garcia, N.A.; et al. Extracellular vesicle secretion is tissue-dependent ex vivo and skeletal muscle myofiber extracellular vesicles reach the circulation in vivo. Am. J. Physiol. Cell Physiol. 2022, 322, C246–C259. [Google Scholar] [CrossRef]
  298. Morel, O.; Jesel, L.; Hugel, B.; Douchet, M.P.; Zupan, M.; Chauvin, M.; Freyssinet, J.M.; Toti, F. Protective effects of vitamin C on endothelium damage and platelet activation during myocardial infarction in patients with sustained generation of circulating microparticles. J. Thromb. Haemost. 2003, 1, 171–177. [Google Scholar] [CrossRef]
  299. Labios, M.; Martinez, M.; Gabriel, F.; Guiral, V.; Munoz, A.; Aznar, J. Effect of eprosartan on cytoplasmic free calcium mobilization, platelet activation, and microparticle formation in hypertension. Am. J. Hypertens 2004, 17, 757–763. [Google Scholar] [CrossRef] [Green Version]
  300. Nomura, S.; Inami, N.; Kimura, Y.; Omoto, S.; Shouzu, A.; Nishikawa, M.; Iwasaka, T. Effect of nifedipine on adiponectin in hypertensive patients with type 2 diabetes mellitus. J. Hum. Hypertens 2007, 21, 38–44. [Google Scholar] [CrossRef] [Green Version]
  301. Suades, R.; Padro, T.; Alonso, R.; Mata, P.; Badimon, L. Lipid-lowering therapy with statins reduces microparticle shedding from endothelium, platelets and inflammatory cells. Thromb. Haemost. 2013, 110, 366–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Lins, L.C.; Franca, C.N.; Fonseca, F.A.; Barbosa, S.P.; Matos, L.N.; Aguirre, A.C.; Bianco, H.T.; do Amaral, J.B.; Izar, M.C. Effects of ezetimibe on endothelial progenitor cells and microparticles in high-risk patients. Cell Biochem. Biophys. 2014, 70, 687–696. [Google Scholar] [CrossRef] [PubMed]
  303. Bulut, D.; Becker, V.; Mugge, A. Acetylsalicylate reduces endothelial and platelet-derived microparticles in patients with coronary artery disease. Can. J. Physiol. Pharmacol. 2011, 89, 239–244. [Google Scholar] [CrossRef] [PubMed]
  304. Morel, O.; Hugel, B.; Jesel, L.; Mallat, Z.; Lanza, F.; Douchet, M.P.; Zupan, M.; Chauvin, M.; Cazenave, J.P.; Tedgui, A.; et al. Circulating procoagulant microparticles and soluble GPV in myocardial infarction treated by primary percutaneous transluminal coronary angioplasty. A possible role for GPIIb-IIIa antagonists. J. Thromb. Haemost. 2004, 2, 1118–1126. [Google Scholar] [CrossRef]
  305. Judge, H.M.; Buckland, R.J.; Sugidachi, A.; Jakubowski, J.A.; Storey, R.F. Relationship between degree of P2Y12 receptor blockade and inhibition of P2Y12-mediated platelet function. Thromb. Haemost. 2010, 103, 1210–1217. [Google Scholar] [CrossRef]
  306. Shouzu, A.; Nomura, S.; Omoto, S.; Hayakawa, T.; Nishikawa, M.; Iwasaka, T. Effect of ticlopidine on monocyte-derived microparticles and activated platelet markers in diabetes mellitus. Clin. Appl. Thromb. Hemost. 2004, 10, 167–173. [Google Scholar] [CrossRef]
  307. Park, S.H.; Belcastro, E.; Hasan, H.; Matsushita, K.; Marchandot, B.; Abbas, M.; Toti, F.; Auger, C.; Jesel, L.; Ohlmann, P.; et al. Angiotensin II-induced upregulation of SGLT1 and 2 contributes to human microparticle-stimulated endothelial senescence and dysfunction: Protective effect of gliflozins. Cardiovasc. Diabetol. 2021, 20, 65. [Google Scholar] [CrossRef]
  308. Akawi, N.; Checa, A.; Antonopoulos, A.S.; Akoumianakis, I.; Daskalaki, E.; Kotanidis, C.P.; Kondo, H.; Lee, K.; Yesilyurt, D.; Badi, I.; et al. Fat-Secreted Ceramides Regulate Vascular Redox State and Influence Outcomes in Patients With Cardiovascular Disease. J. Am. Coll Cardiol. 2021, 77, 2494–2513. [Google Scholar] [CrossRef]
  309. Di Tomo, P.; Lanuti, P.; Di Pietro, N.; Baldassarre, M.P.A.; Marchisio, M.; Pandolfi, A.; Consoli, A.; Formoso, G. Liraglutide mitigates TNF-alpha induced pro-atherogenic changes and microvesicle release in HUVEC from diabetic women. Diabetes Metab Res. Rev. 2017, 33, e2925. [Google Scholar] [CrossRef]
  310. Tardif, J.C.; Kouz, S.; Waters, D.D.; Bertrand, O.F.; Diaz, R.; Maggioni, A.P.; Pinto, F.J.; Ibrahim, R.; Gamra, H.; Kiwan, G.S.; et al. Efficacy and Safety of Low-Dose Colchicine after Myocardial Infarction. N. Engl. J. Med. 2019, 381, 2497–2505. [Google Scholar] [CrossRef]
  311. Nidorf, S.M.; Fiolet, A.T.L.; Mosterd, A.; Eikelboom, J.W.; Schut, A.; Opstal, T.S.J.; The, S.H.K.; Xu, X.F.; Ireland, M.A.; Lenderink, T.; et al. Colchicine in Patients with Chronic Coronary Disease. N. Engl. J. Med. 2020, 383, 1838–1847. [Google Scholar] [CrossRef] [PubMed]
  312. Silvis, M.J.M.; Fiolet, A.T.L.; Opstal, T.S.J.; Dekker, M.; Suquilanda, D.; Zivkovic, M.; Duyvendak, M.; The, S.H.K.; Timmers, L.; Bax, W.A.; et al. Colchicine reduces extracellular vesicle NLRP3 inflammasome protein levels in chronic coronary disease: A LoDoCo2 biomarker substudy. Atherosclerosis 2021, 334, 93–100. [Google Scholar] [CrossRef] [PubMed]
  313. Silva, A.K.A.; Morille, M.; Piffoux, M.; Arumugam, S.; Mauduit, P.; Larghero, J.; Bianchi, A.; Aubertin, K.; Blanc-Brude, O.; Noel, D.; et al. Development of extracellular vesicle-based medicinal products: A position paper of the group “Extracellular Vesicle translatiOn to clinicaL perspectiVEs—EVOLVE France”. Adv. Drug Deliv. Rev. 2021, 179, 114001. [Google Scholar] [CrossRef] [PubMed]
  314. Ma, Q.; Fan, Q.; Han, X.; Dong, Z.; Xu, J.; Bai, J.; Tao, W.; Sun, D.; Wang, C. Platelet-derived extracellular vesicles to target plaque inflammation for effective anti-atherosclerotic therapy. J. Control. Release 2021, 329, 445–453. [Google Scholar] [CrossRef] [PubMed]
  315. Vader, P.; Mol, E.A.; Pasterkamp, G.; Schiffelers, R.M. Extracellular vesicles for drug delivery. Adv. Drug Deliv. Rev. 2016, 106, 148–156. [Google Scholar] [CrossRef] [PubMed]
  316. Yin, M.; Loyer, X.; Boulanger, C.M. Extracellular vesicles as new pharmacological targets to treat atherosclerosis. Eur. J. Pharmacol. 2015, 763, 90–103. [Google Scholar] [CrossRef]
  317. Brill, A.; Dashevsky, O.; Rivo, J.; Gozal, Y.; Varon, D. Platelet-derived microparticles induce angiogenesis and stimulate post-ischemic revascularization. Cardiovasc. Res. 2005, 67, 30–38. [Google Scholar] [CrossRef]
  318. Liu, H.; Gao, W.; Yuan, J.; Wu, C.; Yao, K.; Zhang, L.; Ma, L.; Zhu, J.; Zou, Y.; Ge, J. Exosomes derived from dendritic cells improve cardiac function via activation of CD4(+) T lymphocytes after myocardial infarction. J. Mol. Cell Cardiol. 2016, 91, 123–133. [Google Scholar] [CrossRef]
  319. Wang, C.; Zhang, C.; Liu, L.; Xi, A.; Chen, B.; Li, Y.; Du, J. Macrophage-Derived mir-155-Containing Exosomes Suppress Fibroblast Proliferation and Promote Fibroblast Inflammation during Cardiac Injury. Mol. Ther. 2017, 25, 192–204. [Google Scholar] [CrossRef] [Green Version]
  320. Mayo, J.N.; Bearden, S.E. Driving the Hypoxia-Inducible Pathway in Human Pericytes Promotes Vascular Density in an Exosome-Dependent Manner. Microcirculation 2015, 22, 711–723. [Google Scholar] [CrossRef] [Green Version]
  321. Kervadec, A.; Bellamy, V.; El Harane, N.; Arakelian, L.; Vanneaux, V.; Cacciapuoti, I.; Nemetalla, H.; Perier, 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. 2016, 35, 795–807. [Google Scholar] [CrossRef] [PubMed]
  322. Lima Correa, B.; El Harane, N.; Gomez, I.; Rachid Hocine, H.; Vilar, J.; Desgres, M.; Bellamy, V.; Keirththana, K.; Guillas, C.; Perotto, M.; et al. Extracellular vesicles from human cardiovascular progenitors trigger a reparative immune response in infarcted hearts. Cardiovasc. Res. 2021, 117, 292–307. [Google Scholar] [CrossRef] [PubMed]
  323. Viola, M.; de Jager, S.C.A.; Sluijter, J.P.G. Targeting Inflammation after Myocardial Infarction: A Therapeutic Opportunity for Extracellular Vesicles? Int. J. Mol. Sci. 2021, 22, 7831. [Google Scholar] [CrossRef] [PubMed]
  324. Ratajczak, M.Z.; Ratajczak, J. Horizontal transfer of RNA and proteins between cells by extracellular microvesicles: 14 years later. Clin. Transl. Med. 2016, 5, 7. [Google Scholar] [CrossRef] [Green Version]
  325. Jansen, F.; Yang, X.; Hoelscher, M.; Cattelan, A.; Schmitz, T.; Proebsting, S.; Wenzel, D.; Vosen, S.; Franklin, B.S.; Fleischmann, B.K.; et al. Endothelial microparticle-mediated transfer of MicroRNA-126 promotes vascular endothelial cell repair via SPRED1 and is abrogated in glucose-damaged endothelial microparticles. Circulation 2013, 128, 2026–2038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Akbar, N.; Braithwaite, A.T.; Corr, E.M.; Koelwyn, G.J.; van Solingen, C.; Cochain, C.; Saliba, A.E.; Corbin, A.; Pezzolla, D.; Moller Jorgensen, M.; et al. Rapid neutrophil mobilisation by VCAM-1+ endothelial extracellular vesicles. Cardiovasc. Res. 2022. [Google Scholar] [CrossRef] [PubMed]
  327. Agouni, A.; Mostefai, H.A.; Porro, C.; Carusio, N.; Favre, J.; Richard, V.; Henrion, D.; Martinez, M.C.; Andriantsitohaina, R. Sonic hedgehog carried by microparticles corrects endothelial injury through nitric oxide release. FASEB J. 2007, 21, 2735–2741. [Google Scholar] [CrossRef]
  328. Benameur, T.; Soleti, R.; Porro, C.; Andriantsitohaina, R.; Martinez, M.C. Microparticles carrying Sonic hedgehog favor neovascularization through the activation of nitric oxide pathway in mice. PLoS ONE 2010, 5, e12688. [Google Scholar] [CrossRef] [Green Version]
  329. Comarita, I.K.; Vilcu, A.; Constantin, A.; Procopciuc, A.; Safciuc, F.; Alexandru, N.; Dragan, E.; Nemecz, M.; Filippi, A.; Chitoiu, L.; et al. Therapeutic Potential of Stem Cell-Derived Extracellular Vesicles on Atherosclerosis-Induced Vascular Dysfunction and Its Key Molecular Players. Front. Cell Dev. Biol. 2022, 10, 817180. [Google Scholar] [CrossRef]
  330. Wolf, M.; Poupardin, R.W.; Ebner-Peking, P.; Andrade, A.C.; Blochl, C.; Obermayer, A.; Gomes, F.G.; Vari, B.; Maeding, N.; Eminger, E.; et al. A functional corona around extracellular vesicles enhances angiogenesis, skin regeneration and immunomodulation. J. Extracell Vesicles 2022, 11, e12207. [Google Scholar] [CrossRef]
  331. Saludas, L.; Oliveira, C.C.; Roncal, C.; Ruiz-Villalba, A.; Prosper, F.; Garbayo, E.; Blanco-Prieto, M.J. Extracellular Vesicle-Based Therapeutics for Heart Repair. Nanomaterials 2021, 11, 570. [Google Scholar] [CrossRef] [PubMed]
  332. Wiklander, O.P.B.; Brennan, M.A.; Lotvall, J.; Breakefield, X.O.; El Andaloussi, S. Advances in therapeutic applications of extracellular vesicles. Sci. Transl. Med. 2019, 11. [Google Scholar] [CrossRef] [PubMed]
  333. Sutaria, D.S.; Badawi, M.; Phelps, M.A.; Schmittgen, T.D. Achieving the Promise of Therapeutic Extracellular Vesicles: The Devil is in Details of Therapeutic Loading. Pharm. Res. 2017, 34, 1053–1066. [Google Scholar] [CrossRef] [Green Version]
  334. Robson, A. Exosomes improve myocardial recovery after infarction. Nat. Rev. Cardiol. 2020, 17, 758. [Google Scholar] [CrossRef] [PubMed]
  335. Hofmann, A.; Wenzel, D.; Becher, U.M.; Freitag, D.F.; Klein, A.M.; Eberbeck, D.; Schulte, M.; Zimmermann, K.; Bergemann, C.; Gleich, B.; et al. Combined targeting of lentiviral vectors and positioning of transduced cells by magnetic nanoparticles. Proc. Natl. Acad. Sci. USA 2009, 106, 44–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Vosen, S.; Rieck, S.; Heidsieck, A.; Mykhaylyk, O.; Zimmermann, K.; Bloch, W.; Eberbeck, D.; Plank, C.; Gleich, B.; Pfeifer, A.; et al. Vascular Repair by Circumferential Cell Therapy Using Magnetic Nanoparticles and Tailored Magnets. ACS Nano 2016, 10, 369–376. [Google Scholar] [CrossRef] [PubMed]
  337. Vosen, S.; Rieck, S.; Heidsieck, A.; Mykhaylyk, O.; Zimmermann, K.; Plank, C.; Gleich, B.; Pfeifer, A.; Fleischmann, B.K.; Wenzel, D. Improvement of vascular function by magnetic nanoparticle-assisted circumferential gene transfer into the native endothelium. J. Control. Release 2016, 241, 164–173. [Google Scholar] [CrossRef]
  338. Maiullari, F.; Chirivi, M.; Costantini, M.; Ferretti, A.M.; Recchia, S.; Maiullari, S.; Milan, M.; Presutti, D.; Pace, V.; Raspa, M.; et al. In vivo organized neovascularization induced by 3D bioprinted endothelial-derived extracellular vesicles. Biofabrication 2021, 13. [Google Scholar] [CrossRef]
  339. Chen, C.W.; Wang, L.L.; Zaman, S.; Gordon, J.; Arisi, M.F.; Venkataraman, C.M.; Chung, J.J.; Hung, G.; Gaffey, A.C.; Spruce, L.A.; et al. Sustained release of endothelial progenitor cell-derived extracellular vesicles from shear-thinning hydrogels improves angiogenesis and promotes function after myocardial infarction. Cardiovasc. Res 2018, 114, 1029–1040. [Google Scholar] [CrossRef] [Green Version]
  340. Sekhon, U.D.S.; Swingle, K.; Girish, A.; Luc, N.N.; de la Fuente, M.; Alvikas, J.; Haldeman, S.; Hassoune, A.; Shah, K.; Kim, Y.; et al. Platelet-mimicking procoagulant nanoparticles augment hemostasis in animal models of bleeding. Sci. Transl. Med. 2022, 14, eabb8975. [Google Scholar] [CrossRef]
  341. Ryan, S.T.; Hosseini-Beheshti, E.; Afrose, D.; Ding, X.; Xia, B.; Grau, G.E.; Little, C.B.; McClements, L.; Li, J.J. Extracellular Vesicles from Mesenchymal Stromal Cells for the Treatment of Inflammation-Related Conditions. Int. J. Mol. Sci. 2021, 22, 3023. [Google Scholar] [CrossRef]
  342. Verweij, F.J.; Balaj, L.; Boulanger, C.M.; Carter, D.R.F.; Compeer, E.B.; D’Angelo, G.; El Andaloussi, S.; Goetz, J.G.; Gross, J.C.; Hyenne, V.; et al. The power of imaging to understand extracellular vesicle biology in vivo. Nat. Methods 2021, 18, 1013–1026. [Google Scholar] [CrossRef] [PubMed]
  343. Yadid, M.; Lind, J.U.; Ardona, H.A.M.; Sheehy, S.P.; Dickinson, L.E.; Eweje, F.; Bastings, M.M.C.; Pope, B.; O’Connor, B.B.; Straubhaar, J.R.; et al. Endothelial extracellular vesicles contain protective proteins and rescue ischemia-reperfusion injury in a human heart-on-chip. Sci. Transl. Med. 2020, 12. [Google Scholar] [CrossRef] [PubMed]
  344. Paloschi, V.; Sabater-Lleal, M.; Middelkamp, H.; Vivas, A.; Johansson, S.; van der Meer, A.; Tenje, M.; Maegdefessel, L. Organ-on-a-chip technology: A novel approach to investigate cardiovascular diseases. Cardiovasc. Res. 2021, 117, 2742–2754. [Google Scholar] [CrossRef] [PubMed]
  345. Verweij, F.J.; Hyenne, V.; Van Niel, G.; Goetz, J.G. Extracellular Vesicles: Catching the Light in Zebrafish. Trends Cell Biol. 2019, 29, 770–776. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic summary of extracellular vesicles (EVs) as therapeutic tools. EV indicates extracellular vesicle; MI, myocardial infarction; ↑, increase; ↓ decrease.
Figure 1. Schematic summary of extracellular vesicles (EVs) as therapeutic tools. EV indicates extracellular vesicle; MI, myocardial infarction; ↑, increase; ↓ decrease.
Cells 11 01845 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Suades, R.; Greco, M.F.; Padró, T.; Badimon, L. Extracellular Vesicles as Drivers of Immunoinflammation in Atherothrombosis. Cells 2022, 11, 1845. https://doi.org/10.3390/cells11111845

AMA Style

Suades R, Greco MF, Padró T, Badimon L. Extracellular Vesicles as Drivers of Immunoinflammation in Atherothrombosis. Cells. 2022; 11(11):1845. https://doi.org/10.3390/cells11111845

Chicago/Turabian Style

Suades, Rosa, Maria Francesca Greco, Teresa Padró, and Lina Badimon. 2022. "Extracellular Vesicles as Drivers of Immunoinflammation in Atherothrombosis" Cells 11, no. 11: 1845. https://doi.org/10.3390/cells11111845

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