Skip to main content

Circulating microparticles: square the circle

Abstract

Background

The present review summarizes current knowledge about microparticles (MPs) and provides a systematic overview of last 20 years of research on circulating MPs, with particular focus on their clinical relevance.

Results

MPs are a heterogeneous population of cell-derived vesicles, with sizes ranging between 50 and 1000 nm. MPs are capable of transferring peptides, proteins, lipid components, microRNA, mRNA, and DNA from one cell to another without direct cell-to-cell contact. Growing evidence suggests that MPs present in peripheral blood and body fluids contribute to the development and progression of cancer, and are of pathophysiological relevance for autoimmune, inflammatory, infectious, cardiovascular, hematological, and other diseases. MPs have large diagnostic potential as biomarkers; however, due to current technological limitations in purification of MPs and an absence of standardized methods of MP detection, challenges remain in validating the potential of MPs as a non-invasive and early diagnostic platform.

Conclusions

Improvements in the effective deciphering of MP molecular signatures will be critical not only for diagnostics, but also for the evaluation of treatment regimens and predicting disease outcomes.

Background

The present review summarizes information concerning microparticles (MPs), covering the clinical aspects of circulating MPs, recent advances and technological developments in this field.

Implementation

Several recent reviews have concentrated on specific aspects of cellular vesicles biology, focusing primarily on exosomes (subset of cellular vesicles with size < 100 nm) and the mechanisms involved in cellular vesicles release and signaling [16]. This review focuses on another subset of cellular vesicles, i.e. microparticles (MPs). MPs are submicron vesicular fragments of cells that can be released by diverse eucaryotic and procaryotic cells and multicellular organisms under conditions of stress/injury [79]. Although novel methods to identify and characterize MPs have been developed in the last decade, classification of MPs, understanding of the molecular mechanisms of their release and biological function are still under intensive scrutiny [1014]. The aims of this review article are to provide i) a systematic overview on circulating MP biology, and ii) a comprehensive description of the role of MPs in different diseases, based on the analysis of over 200 publications addressing changes in circulating MPs during pathological processes.

Results and discussion

MPs: attempts to define

MPs are described as a heterogeneous population of membrane-delimitated vesicles 50–1000 nm in size released from the cells in which they form and retaining certain antigens of their cells of origin [8, 14, 15]. MPs could be distinguished from other groups of cell-derived vesicles such as exosomes and apoptotic bodies. Exosomes are small vesicles (40–100 nm) that form through constitutive exocytosis of multivesicular endosomes [4, 8], and often contain endocytic markers, such as tetraspannins and HSP73 [2, 16]. MPs (also called “ectosomes”) form mostly by reverse budding and fission of the plasma membrane [17]. Because exosomes and MPs are often released concomitantly, differentiation of these two microvesicular species is difficult [18].

The size of MPs (50 to 1000 nm), their lipid composition, and their irregular shape and density are major parameters that separate them from exosomes (usually of diameter < 100 nm and lower density – 1.13-1.19 g/mL) and apoptotic bodies (much larger vesicles released at the final steps of apoptosis and normally 1000–3000 nm in size) [8, 19]. This variance in reported size of MPs could occur due to limitations in the methods of the detection of MPs and differences in MP purification protocols, such as the anticoagulant used, centrifugation speed, filtration conditions, and type of storage used [20, 21]. Besides, the majority of MPs express on their surface phosphatidylserine (PS) whereas PS is usually absent from exosomes’ surface [22]. In general, exosomes are smaller than MPs; however, reported sizes of MPs vary by publication, ranging from 50 nm to 1000–2000 nm (Additional file 1) and thus it is better to say that different research protocols allows one to enrich preparation with certain type of vesicles but not to separate them as a pure fraction. Current nomenclature of cell-derived vesicles was exhaustively presented recently [8], and we will follow it using terms microparticle and microvesicle as synonyms.

Methods of MPs characterization

Isolation of MPs typically involves a combination of centrifugation and size-based filtration followed by characterization using flow cytometry, electron microscopy, Western blotting or proteomics. Isolation of MPs from the peripheral blood of patients or healthy controls starts with drawing blood into the tubes with different anticoagulants: sodium citrate, acid-citrate-dextrose, EDTA salt, or heparin. Sodium citrate is the most widely used anticoagulant [23]; however, blood collected with sodium citrate usually gives significantly lower levels of PS-positive MPs than blood collected in heparin [24]. Centrifugation is a critical step as well, since it can induce additional shedding of MPs from some cell types [2426]. It is also possible that MPs can fuse during preparation, as MPs isolated by centrifugation are somewhat bigger than MPs in native MP-containing biological samples [27]. Haemolysis during sample preparation can significantly affect the amount of MPs isolated from plasma, as well as amounts of MP-related molecules like miRNA [28]. The size distributions of platelets (2–3 μm) and MPs (up to 2 μm) partially overlap, and current consensus indicates that the best way to remove contaminating platelets from MP preparations is via filtration. However, filtration of MPs should be used with caution, since this procedure can lead to fragmentation of larger MPs [29]. Finally, storage of purified MPs even at −80°C may further modify their characteristics [24, 30].

Research focused on elucidating MP composition and functional activity is hampered by the complexity of the biological fluids where MPs are present and the small size of MPs [31]. Electron microscopy (EM) gives the diameter of individual MPs, but does not always provide quantitative data on the MP population - particularly when negative staining or cryoelectron microscopy are used. On ultrathin sections MPs appear as single, membrane-bounded vesicles with diameters ranging between 20–40 nm [3235] and 300–700 nm [3642], with the larger MPs exhibiting heterogeneous internal content. MPs as large as 1 μm in diameter were described using freeze-fracture and scanning EM [3235]. Besides EM, atomic force microscopy and dynamic light scattering have been used for MP characterization [21, 27, 29, 31].

The protein content of MPs is usually ascertained by Western blotting and proteomic approaches [43, 44]. These assays require large numbers of MPs, limiting their utility for translational studies that require serum or other bodily fluids [45]. To date, only flow cytometry and microscopy methods have proved capable of providing specific information on the presence or absence of specific antigens in MPs derived from limited amounts of material. The application of different methods to exosome and MP research has been summarized by Van der Pol and coauthors [8, 31, 46], and in a number of recent publications [2124, 47, 48].

Current flow cytometry methods utilize both fluorescence probes and light scattering. Quantification of MPs by flow cytometry shows good correlation with the relative light scattering intensities determined by dynamic light scattering [49]. There are also indirect approaches for MP enumeration based on their functional activities [50, 51]. However, conventional flow cytometry light scattering has size limitations and usually not able to detect microvesicles with diameters smaller than 300–400 nm as a separate fraction [31, 52]. Particle size can be directly measured using impedance-based Coulter-type cytometers, but the sensitivity of this technology is also limited by 300–500 nm [31, 52, 53]. One other widely employed cytometric approach for the identification and characterization of MPs involves the use different sized beads as references [53, 54]. However, the refractory index of polystyrene or other synthetic beads is higher than that of MPs, thus signals generated by MPs are very small. While conventional cytometers equipped with a photodiode for measuring forward light scatter have significant limitations in sensitivity for MP analysis, cytometers equipped with a photomultiplier in the forward scatter channel allow for better resolution of MP fractions (Figure 1, SORP FACSAria (BD Biosciences, San Jose, USA)). MPs can be directly stained with fluorescent antibodies and with fluorescent lipophilic dyes, both of which dramatically increase the ability of the cytometer to separate MPs from debris. For the best detection, MP staining for flow cytometry should include a lipid marker such as calcein AM, PKH67, or bio-maleimide [5456], since staining MPs with only specific antibodies (AB) or annexin V can leave a significant percentage of MPs unstained or poorly stained and, as a result, lead to underestimation of MP levels. Recently, investigators have begun to use flow image cytometry for MP characterization (Figure 2). The advantages and disadvantages of commonly used methods for MP quantification and characterization are summarized in Table 1.

Figure 1
figure 1

Distribution of Dragon Green-conjugated beads in sizes of 190 nm, 510 nm, and 730 nm (images acquired with a SORP Aria 2 cytometer, Flow and Imaging Cytometry Resource, PCMM, Boston Children’s Hospital). This figure was included in an advanced abstract as part of the Proceedings of the International Workshop on Applied Cytometry (2012).

Figure 2
figure 2

MP images (erythrocyte-derived MPs) acquired with an Imagestream 100 (40x objective) (Amnis Inc, Seattle, USA). A dotblot showing a mixture of erythrocytes and erythrocyte-derived microparticles (X-axis-brightfield area; Y-axis-brightfield aspect ratio of intensity). a.) Multiple erythrocytes (region R3 on dotblot); b.) Single erythrocytes (region R2 on dotblot); c.) Microparticles (brightfield) (region R1 on dotblot); d.) Microparticles stained with calcein AM (images from calcein AM-channel are particles taken from R1-region).

Table 1 Summary of some methods applied for MPs research

Origin of MPs

MPs have been identified in human plasma, peripheral blood, cord blood, urine, saliva and cerebrospinal fluid [45, 5762]. In addition, MPs have been found at different sites in lung disease patients, such as in the sputum from cystic fibrosis patients [39], and in bronchoalveolar lavage fluid (BALF) from patients with acute respiratory distress syndrome or hydrostatic pulmonary edema [63, 64]. MPs have also been described in human atherosclerotic plaque [6567], ascites, postoperative drainage fluid, and chyloid fluid [41], as well as in immunologically privileged sites such as vitreous eye liquid and synovial liquid [6872]. Large body of evidence suggests that MPs are derived from all cellular types. The origin of MPs is critical because MPs with similar shapes and diameters yet derived from different cell types possess unique functional capabilities. Aleman et al. showed that MPs (100–300 nm in size) derived from monocytes had higher ability to support clot formation, making it more dense and stable compared to PMPs [73]. It has long been thought that the majority of MPs in the peripheral blood of a healthy person are released from platelets and endothelial cells [24, 74]. However, it was recently suggested that CD61-positive MPs (currently called “PMPs”) originate directly from megakaryocytes [75, 76]. Rank et al. showed that patients undergoing hematopoietic stem cell transplantation after total body irradiation (12 Gy) exhibit a rapid decline of the level of peripheral blood MPs, with CD61+ MPs disappearing faster than platelets and MPs expressing CD63 or P-selectin, leading the authors to conclude that at least a fraction of CD61+ MPs originate from megakaryocytes [77].

To characterize the cellular origin of MPs in peripheral blood, the most common approach is to stain MPs with fluorescently-labeled AB directed against antigens of parental cells (for example CD41, CD61 and platelet-activation marker CD62 for platelets; glycophorin for erythrocytes; CD45 for lymphocytes; CD14 for monocytes; and so on) and to perform subsequent analysis by flow cytometry. However, a large variety of CD markers have been used by different groups to characterize background and activation of MPs derived from endothelial cells (CD31, CD34, CD62E, CD51, CD105, CD144, CD146) versus platelets (CD41, CD41a, CD42a, CD42b, CD61, CD62P) may have led to inconsistency in the functional characterization of MPs populations (reviewed in [15]).

Shedding (ectocytosis) and MP content

Though MP shedding is enhanced upon cell activation, constitutive ectocytosis is a permanent ongoing process in vivo for the majority of cells and significant levels of MPs originating from different cells can be always found in the plasma [78, 79]. MPs contain a wide range of biomolecules: proteins (signal proteins and receptors, cytoskeleton and effector proteins), lipids, and nucleic acids, (e.g. microRNA, mRNA, and even DNA). MP surface protein content may be different from that of the plasma membrane of the cell of origin, as the incorporation of protein molecules into MPs can be a selective and modulated by agonist activators and/or microenvironments of the parental cells [54, 8084]. Depending on the stimulus, the protein content of MPs derived from the same cell lineage can vary. Jimenez et al. [85] demonstrated that endothelial cells release qualitatively and quantitatively distinct MPs in response to TNF-α (activation stimulus) and upon the induction of apoptosis by growth factor deprivation. In addition, several groups performing MP proteomic profile studies have found that characteristics of MPs isolated from peripheral blood depend on the type of stimulus used for their generation [54, 86]. It has been shown that the density of β2-integrin and P-selectin is markedly enhanced in platelet-derived MPs (PMPs), whereas MPs from activated neutrophils are highly enriched in activated Mac-1 (10-fold enrichment) [87, 88]. Moreover, the surface of PMPs is 50 to 100-fold more procoagulant than the surface of activated platelets [87]. It is likely that specific protein enrichment of MPs membrane is due, at least in part, to lateral re-organization of membrane lipids into cholesterol-rich lipid rafts during MP shedding [89, 90]; however, the exact mechanisms involved in this process requires further investigation.

Plasma membrane remodelling is a critical event during apoptosis and cell activation, and enzymes that regulate this process also regulate MP production [14]. The formation of MPs in response to activating stimuli is initiated by an agonist-mediated increase in intracellular calcium (Figure 3a), activation of kinases and inhibition of phosphatases, and calpain activation [14]. Activation of calcium-dependent scramblase (an ATP-independent transporter) and floppase (an exofacially-directed, ATP-dependent transporter) [91] results in exposure of PS on the outer leaflet of the plasma membrane [92]. Levels of PS exposure depend on the type of stimulation [85, 9395]. However, in some cases the processes of PS exposure and MP generation can be separated [96]. Particularly in endotoxemia and sickle cell disease formation of a large number of annexin-negative MPs was described [97, 98]. Concomitant with the exposure of PS on the outer leaflets of MP membranes, calcium-sensitive enzymes such as calpain and gelsolin are activated, which promotes subsequent vesiculation [99]. In addition to the pathways decribed above, MP formation and trafficking can occur via ARF6-regulated endosomal pathways [100]. The exact mechanisms of lipid scrambling, PS exposure on the outer membrane leaflet, and ultimately MP formation, can differ between cell types [101, 102]. In any case, PS on the surface of MPs is an important factor in mediating their functional activity: PS acts as a major prothrombotic and procoagulation signal, enhancing activation of coagulation proteins, TF, and platelet aggregation [103]. The functional role of PS-negative MPs is still a subject of debate, though elevated levels of circulating Annexin-negative MPs had been reported for initial phase of stroke, systemic lupus erythematosus (SLE) and some other diseases [104107]. MPs can be captured by PS-binding molecules like T-cell immunoglobulin domain and mucin domain proteins, which are expressed on the surface of activated lymphocytes and phagocytes [108, 109]. Formation and/or release of MPs can also be influenced by apoptotic signals [110] (Figure 3b). The shedding of MPs in response to apoptotic stimuli critically depends on the activation of Rho-associated kinase ROCK1 [111].

Figure 3
figure 3

Basic scheme of MP formation via reverse budding. A. Activated cells release MPs in response to Ca++ agonists. Increased concentration of Ca++ alter the asymmetric PS distribution of the plasma membrane, activate kinases, inhibit phosphatases and activate calpain, which leads to reorganization of cytoskeleton and increased MPs production. B. MP formation during the early stages of apoptosis is associated with GTP-bound Rho proteins, which activate the ROCK-I kinase. This kinase is involved in cortical myosin-II contraction, detachment of the plasma membrane from the cytoskeleton, and release of MPs that have hijacked cytoplasmic components, nucleic acids, and membrane antigens.

Several other enzymes possibly involved in MPs formation and activity include aminophospholipid translocase, and other members of the floppase family, as well as protein disulfide isomerase and acid sphingomyelinase [58, 112114]. Protein disulfide isomerase (PDI) – enzyme modulating flippase and floppase activities and regulating coagulation on endothelial cells [112] was shown to be a component of MPs released during tissue factor (TF)-dependent thrombosis [113]. Recently, Bianco et al. [114] demonstrated that activation of acid sphingomyelinase is necessary and sufficient for MP release by glial cells. As mentioned above, it is likely that lipid rafts are important participants in MP formation, since the depletion of plasma membrane cholesterol or raft disruption by methyl-cyclodextrin reduces MP release from a variety of cell types [89, 115, 116].

Enhanced release of MPs is associated with diverse stimuli including hormones, fatty acids, reactive oxygen species (e.g. hydrogen peroxide) [117], increased intracellular calcium levels [99]. Increased MP output is also driven by signals transduced through specific activating receptors, such as the purinergic receptor P2X on monocytes and neutrophils, thrombin receptors on platelets, and Toll-like receptor 4 (TLR4) on dendritic cells [118]. The level of MPs in human plasma can increase or decrease in response to different hormones, such as progesterone, estradiol, estrogen, insulin and others [119121]. For example, low levels of estrogen in the blood are associated with increased microvesiculation and MP release [122]. Treatment with glucocorticoids significantly decreases the level of PMPs in peripheral blood in patients with polymyositis or dermatomyositis [123]. While insulin may promote MP release in certain cases, it has been found to reduce the procoagulant activity of MPs derived from lipopolysaccharide (LPS)-activated monocytes [124].

MPs also carry all types of nucleic acid molecules, including mRNA and DNA fragments [125, 126]. Risitano et al. [127] demonstrated that platelet-derived mRNA could be transferred by MPs to monocytic and endothelial cell lines and undergo translation in the recipient cells. Improved ability to detect low copy numbers of small RNAs, including miRNA, has rapidly advanced the MP field, since these molecules has to be porotected from plasma nucleases and may be functional only when had been transferred by MPs internalized by target cells. Indeed, MPs from healthy donors contain miRNAs that have different functional activities [128], such as regulation of hemostasis [129]. Diehl and coauthors [130] assessed miRNA profiles of MPs derived from stimulated and non-stimulated endothelial cells (THP-1 and HUVECs) and found that miRNA profiles of MPs differed from those found in the stimulated or non-stimulated parental cells (some miRNAs upregulated while others down-regulated), suggesting a process of selective miRNA packaging into MPs. Specifically, MPs derived from stimulated THP-1 cells contained increased inflammatory miRNA and induced inflammation markers up-regulation in non-stimulated cells [130].

Functional activities of MPs: interaction with homologous or heterologous cells

As outlined above, MP production is a tightly regulated and selective process, suggesting that MPs may be important mediators of cell-to-cell communication. MPs can be internalized in a dose-dependent manner by macrophages, endothelial cells and other cell types (an example of MP internalization by hCMEC/D3 cells is shown in Figure 4). MP internalization can influence both functional and phenotypic characteristics of target cells. MPs may operate via surface interactions with receptor molecules on target cells or, more importantly, by directly transferring their contents, including RNA [130133], bioactive lipids (for example platelet-activating factor (PAF) and PAF-like lipids), and proteins into the recipient cell [134, 135].

Figure 4
figure 4

An epifluorescence microscopy image shows that hCMEC/D3 cells have internalized small MPs (arrowheads), which had been purified from human glioma cells treated with TRAIL (100 units/ml) and stained with PKH-67 (Sigma, USA) before addition to the hCMEC/D3 cells. MPs that are attached to the cell surface are out of focus (representative photo from Z-stack collection). Objective Plan Apo x60/1.4. Bar 5 μm.

The MPs express adhesion molecules on their surface, which may influence the probability of their capture by target cells and mediate MPs effects on cell behavior [136138]. The cellular origin and site of release are essential factors in determining the functional activities of MPs. For example, MPs derived from red blood cells, but not from blood polymorphonuclears (PMNs) inhibit activation of macrophages by zymosan and LPS [139, 140]. MPs participate in the release of insoluble proteins such as transmembrane receptors (CCR5, TF, EGFR, etc.) [90, 141, 142] and other surface molecules involved in immunomodulation [118, 143, 144]. The transfer of membrane-anchored receptors by MPs results in phenotypic alteration of the recipient cell, making it susceptible to different activating stimuli. For example, transfer of the chemokine receptor CCR5 by MPs to CCR5-deficient peripheral blood mononuclear cells makes them more sensitive to infection by CCR5-tropic HIV viruses [141]. Shuttling of the chemokine receptor CXCR4 by MPs contributes to HIV disease progression, since CXCR4 also serves as a co-receptor for some viruses [145]. Besides transferring receptor molecules, MPs may transfer chemokines, cytokines and growth factors to target cells [90, 146]. For example, MPs transfer pro-apoptotic arachidonic acid between endothelial cells and circulating angiogenic cells [147], and constitute a main reservoir of blood-originated TF, the main activator of blood coagulation [142].

Lung-derived MPs have been shown to transfer mRNA to marrow cells [148], and MPs derived from endothelial progenitor cells have been reported to carry a wide range of mRNAs and to promote angiogenic activity and proliferation in quiescent endothelial cells [149]. Hemopoeitic stem cell-derived MPs contain mRNAs that contribute to the reprogramming of target cells [150]. Transfer of mRNAs to hepatocytes by liver stem cell-derived MPs induce proliferation and resistance to apoptosis [151]. Yuan et al. [152] demonstrated that miRNAs that are highly enriched within MPs are transferred to mTEC cells via MP internalization. miRNAs shuttled by MPs have been shown to downregulate the activity of proteins participating in cell proliferation and apoptosis such as cyclin D1, Bcl-2 and PTEN [153]. The most abundantly expressed miRNA in plasma MPs is miR-223, which participates in the maturation, proliferation and differentiation of myeloid and lymphoid cells [128]. MPs may also assist in the delivery to target cells of synthetic miRNAs [153].

A growing body of evidence supports an important role for MPs in the induction of apoptosis. MPs released at the early stages of apoptosis do not contain organelles and their size is smaller than 1 μm; however, they sediment at a lower acceleration than exosomes [110]. In contrast, so-called “apoptotic bodies”, which are released during the final stages of apoptosis, have a size of 1–4 μm, and often contain organelles [144]. Recently, Sarkar et al. [154] have demonstrated that monocyte-derived MPs induce death of target cells by delivering caspase-1. MPs from endothelial cells and platelets may also contain active executive caspase-3 [155157]. Similarly, tumor-derived MPs serve as circulating cargoes for Fas ligand (FasL or CD95L), and therefore induce apoptosis in lymphoid target cells harboring the Fas receptor [158, 159]. In addition to FasL, MPs and exosomes from different human tumors (melanoma, head, neck, ovary, colorectal and other cancers) may carry other proapoptotic molecules, such as TRAIL [143, 159161].

Circulating MPs

The level of circulating MPs depends on the balance between their rates of formation and clearance. Clearance of MPs occurs through several main mechanisms. The major one is degradation due to the action of phospholipases and proteases [162]. Other potential routes of MP clearance include: (i) opsonization with subsequent phagocytosis; (ii) uptake of MPs from the circulation by liver Kupffer cells in a PS-dependent manner [163]; (iii) phagocytosis of MPs by splenocytes [164]; and (iv) uptake of MPs by the lung macrophages [165]. In a rat model, both the spleen and liver were found to participate in the clearance of MPs labeled with radioactive 51Cr, with only 12% of injected erythrocyte-derived microvesicles retained in the plasma after 60 min [166]. However, recent studies suggest that survival of PS+ MPs in human blood is rather long: the half-life of Annexin V+-MPs measured upon transfusion of apheresis platelet concentrates is approximately 5.8 hours and for CD61+ MPs it is 5.3 hours [167]. MP size is also an important factor in their clearance – strong inverse correlation between IgM-mediated clearance half-time and particle size of MPs by macrophages was determined [168]. On opposite, Al-Faraj et al. [169] demonstrated rapid clearance (within 5 min) of iron-labeled MPs by time-lapse molecular imaging using mouse model. However, it should be taken into account that labeling of such a fragile thing as MPs ex vivo may change clearance characteristics and kinetics.

While low MP concentrations can be detected in the blood and body fluids of healthy subjects [170172] (summarized at Table 2), increased concentrations of MPs in the blood of patients with different pathological states supports the notion that MPs play a role in numerous diseases, including different cancers (Table 3), infectious diseases, autoimmune diseases, thromboembolic events and others (Table 4). However, most of these studies are observational and the possible role of MPs as prognostic biomarkers in stratification of disease risk groups is only starting to be addressed. There have been very few prospective studies aimed at evaluating whether there is an association between the quantities of a certain subtype of MP (endothelial, erythrocyte or other cell-derived MPs) and the outcome of diseases or therapeutic procedures [173175]. Increased MP levels in pathological disorders such as intracerebral hemorrhage, endotoxemia, hepatitis C and others are generally associated with adverse outcomes (Additional file 3), and high levels of MPs associated with these disorders could, at least partly, be implicated in the vascular complications of these diseases. However, although increased levels of circulating MPs have been associated with various autoimmune diseases (SLE, rheumatoid arthritis, systemic sclerosis), facile correlation of MP quantity and adverse outcomes is complicated by the fact that plasma MP levels appear to increase to lower levels in patients with more severe disease [176]. Thus, the factors regulating MP release during desease progression are complex and yet remain to be evaluated. In this regard, it is important to consider the effect of pharmacological agents on circulating MP levels and their composition (summarized in the Additional file 4). Most of these studies have demonstrated that beneficial treatment of disease lowers circulating MP levels. For example, treatment of multiple sclerosis (MS) with interferon-β1 decreased the amount of circulating CD31+ endothelial MPs in plasma [177]. Similar results were obtained by Lowery-Nordberg et al. [178]. These data suggest that the quantity of specific MPs in the circulation may be used as a surrogate marker for interferon therapy responsiveness.

Table 2 MPs levels in the plasma of healthy controls
Table 3 MPs levels in the plasma and body fluids of patients with cancer
Table 4 MPs levels in the plasma and body fluids of patients with different disorders

The association of elevated levels of certain MP subtypes with specific disease states may also have therapeutic implications. An interesting possibility is the use of in vitro generated MPs to stimulate neovascularization in the diseases with impaired angiogenesis [179], while a different subset of MPs could be used to inhibit tumor-induced angiogenesis and, possibly, even tumor development [180]. Therapeutic strategies to reduce severity of disease may also decrease the level of circulating MPs. Thus, the level of platelet-derived MPs in diabetic patients is decreased after treatment with antioxidants such as vitamin C [181] or miglitol [182]. La Vignera et al. [183] showed that endothelial-derived MP (EMPs) level is significantly decreased in patients with erectile dysfunction after treatment with tadalafil. The concentration of erythrocyte-derived MPs (ErMPs) in patient blood correlates with severity of malaria disease and starts to decrease 24 hours after the beginning of antimalarial treatment, reaching baseline values after two weeks of treatment in patients infected with P.vivax and P.malariae, but after more prolonged therapy in patients with P.falciparum [184].

These findings have ignited interest to MPs as possible biomarkers for diagnostics and evaluation of efficiency of a therapeutic strategy.

MPs in cancer

Cancer cell-derived MPs have been studied intensively in recent years, and their potential as diagnostic and prognostic tools has been described [185, 186]. Tumor-derived MPs carry specific molecular markers typical for the cells of their origin, including epithelial cell adhesion molecule (EpCam), human epidermal growth receptor 2 (HER-2), CCR6, extracellular metalloproteinases (MMPs), vascular endothelial growth factor (VEGF), and some others [118, 187191]. However, many types of cancer, such as ovarian and pancreas malignancies, exhibit no specific biomarker that makes their screening or early detection difficult. Several groups have described the transfer of oncogenic proteins and chemokines between cells by tumor-derived MPs, which leads to the horizontal spread of aggressive phenotypes among tumor cells had not expressing these proteins by themselves [90, 192]. MPs from cancer cells contain a variety of cell-surface receptors, cytoskeletal components and intracellular signaling proteins [192] and the concentration of tumor-derived MPs increases during tumor progression [186, 189]. Peripheral blood from cancer patients contains not only cancer cell-derived MPs but also high levels of procoagulant and platelet-derived MPs [190], which may contribute to the development of clinically relevant haemostatic abnormalities in cancer patients that is referred to as Trousseau’s syndrome [193]. Reprogramming of target cells by MPs was first described by Ratajczak et al. [122], and later on it has been shown directly that exposure of normal cells to cancer cell-derived MPs that contain fibronectin and tissue transglutaminase causes the recipient cells to acquire a transformed phenotype [194]. Moreover, it was reported that when MPs produced by cultures of different human primary tumors or established tumor cell lines were isolated and added back to the same cancer cells the growth of these cells was accelerated [90]. Finally, it was found that MPs derived from a subset of CD105+ tumor-initiating human renal carcinoma cells were able to activate endothelial cells in vitro and triggered their growth and vascularization after implantation into SCID mice [195].

MPs shed by tumor cells serve as a profound additional pathway for drug release [196]. Intensity of MP shedding and anti-cancer drug resistance by positively correlate across wide number of cell lines and drugs tested [196]. Besides, Jaiswal et al. [197] have shown that MPs derived from both ABCB1-mediated multidrug-resistant acute lymphoblastic leukemic and breast cancer cells can transfer mRNAs that encode multidrug resistance (MDR) transporter proteins into the drug-sensitive cancer cells, allowing for horizontal acquisition of drug resistance. This study also demonstrated that MPs express greater concentration of specific miRNAs as compared to their cells of origin (for example miR-451). This “non-genetic” intercellular transfer of molecular components provides an alternative pathway for circumvention of MDR. The time-dependence of P-gp transfer by MPs and increase of influx activity in MCF-7 breast cancer cells reveal the occurence of multiple routes for extragenetic MDR acquisition by cancer cells [198].

The contribution of platelet-derived MPs to hematogeneous cancer metastasis is tied to their procoagulant activity [199]. Metastatic processes depend on the haemostatic competence of tumour cells and their capacity to initiate microvascular thrombosis [190], and MPs may promote these processes via transfer of mRNAs that encode angiogenic factors such as MMP-9, interleukin-8, VEGF [200]. Indeed, injection PMP-covered Lewis lung carcinoma cells (LLC) into syngeneic mice results in the formation of significantly more metastatic foci in the lungs of these animals as compared to mice injected only with LLC [200]. Also in prostate cancer patients elevated plasma PMP levels correlate with aggressiveness of tumors and poor clinical outcome [201].

MPs and vascular diseases

Platelet-derived MPs have been extensively investigated for their ability to induce coagulation and participate in thrombosis because they display PS and other negatively charged phospholipids that provide binding sites for activated coagulation factors [202]. PMPs have significantly higher (50-100x) procoagulant activity compared even to activated platelets [87]. PMPs may regulate additional vascular pathways, including activation of endothelial cells and leukocytes, stimulation of angiogenesis, and induction of apoptosis in endothelial cells [203]. MPs released by normal endothelial cells are implicated in angiogenesis, as well as bone regeneration and mineralization in vivo [204206]. MPs originating from human atherosclerotic plaques carry mature form of tumor necrosis factor (TNF)-converting enzyme metalloprotease TACE/ADAM 17, which cleaves TNF and its receptors TNF-R1 and TNF-R2 [207]. These MPs enhance shedding of TNF from cultured human cells that overexpress TNF, as well as TNFR1 shedding from HUVEC cell lines, suggesting that TACE+ MPs regulate the inflammatory balance in culprit atherosclerotic plaque lesion [207]. Several forms of hemolytic anemia are associated with elevated levels of MPs in plasma and concomitantly with high tissue factor (TF) activity [97, 208210]. Monocyte-derived MP levels are elevated in the plasma of paroxysmal nocturnal hemoglobinuria patients, as monocytes in these indviduals are fragile due to a deficiency in surface expression of CD55 and CD59 [209].

Since endothelial MPs from patients with metabolic disorders induce endothelial dysfunction in animal models [211], and elevated circulating MP levels are associated with both severity and adverse outcomes in several cardiovascular pathologies, including myocardial infarction, atherothrombosis, hypertension, and preeclampsia, risk stratification for these conditions now relies, in part, on the measurement of MP levels (summarized in Additional file 3).

MPs and infectious diseases

Bacterial virulence factors such as the M1 protein from S.pyogenes and lipopolysaccharide (LPS) from E.coli stimulate the release of procoagulant MPs from PBMCs [212, 213]. A number of publications have reported that specific MP subtypes in septic patients, such as endothelium-, platelet- and monocyte-derived MPs, are associated with different etiologies of sepsis (S.pyogenes, Staphylococcus, Pneumococcus, Enterococcus) [213, 214]. Elevated MP levels are associated with systemic inflammatory response syndrome (SIRS) and hemolytic uremic syndrome caused by E.coli infection [215, 216]. It is possible that MPs produced by infected cells, or by cells exposed to bacterial virulence factors, may contribute to secondary organ dysfunction observed during these disorders. Mastronardi and colleagues [217] have reported that injection of MPs from septic shock patients into experimental animals leads to changes in the enzyme systems related to inflammation, nitrative and oxidative stress. These findings are in accordance with the results obtained by other investigators [218], which have indicated that the injection of normal rats with MPs obtained from septic rats induces hemodynamic changes and septic inflammatory responses in the heart.

ErMP levels are significantly increased in the blood of malaria patients with coma or severe malaria [184] and correlate with plasma TNF concentrations [219]. Cell-derived and Plasmodium-derived MPs contribute to the development of fatal cerebral malaria [220222]. In in vitro experiments PMPs were found to bind preferentially to Plasmodium-infected erythrocytes or iRBCs, and increase cytoadherence of iRBCs to HUVECs [222]. Moreover, it has been shown that P.falciparum synthesizes and packages Maurer’s clefts* (*parasite-derived structures within the host cell cytoplasm that are thought to function as a sorting compartment between the parasite and the parasitophorous membrane [223]) subsequently exporting them to the cytoplasm of infected erythrocytes via MPs shedding [223]. Observations on another eukaryotic parasite, L.donovani, also demonstrated that parasite-produced microvesicles are released from infected cells [224]. MPs released by bacteria Porphyromonas gingivalis that cause periodontitis disease, carry lipoproteins and other proinflammatory mediators to the distant sites and contribute to progression of atherosclerosis [225, 226]. Summarizing it could be concluded that in many cases MPs and exosomes released by infected host cells contain pathogen-derived antigens and virulence factors and may modulate disease progression and immune response [225230].

Conclusion

As methods for isolating and characterizating MPs advance, it is anticipated better understanding of the mechanisms of MP formation and functional activity will be achieved in near future (a current overview of MP activity is summarized in Figure 5). Flow cytometry, fluorescent microscopy and light scattering methods will be critical for the characterization of MP preparations. A growing number of reports have demonstrated that MPs are produced by a remarkably diverse array of cell types and may alter the phenotype and behavior of different cell populations. However, despite four decades of MP research, we are just beginning to understand the contribution of MPs to disease development and pathogenesis. The association of elevated MP levels with many different pathological states makes them of particular interest for clinical research, and suggests that these tiny vesicles have great potential for the development of new diagnostic assays aimed at identifying early stages of pathological disorders and response for therapy, the creation of a novel class of therapeutics for improved intervention in a group of difficult-to-treat diseases. Future diagnostic exploitation of MPs may circumvent the need for some current invasive procedures, such as amnioscentesis or chorion villus sampling for the diagnosis of prenatal disorders. Further dissection of circulating MP components and their functional roles will undoubtly expand their usefulness as biomarkers and, in turn, as sentinels that steer investigators to more efficacious treatment options.

Figure 5
figure 5

Potential mechanisms of MP action.

Abbreviations

AB:

Antibody

ABCA1:

ATP binding cassete transporter A1

ABCB1:

ATP binding cassete transporter B1

ADAM10:

A disintegrin and metalloproteinase domain-containing protein 10

ARF6:

ADP-ribosylation factor 6

Allo-HSCT:

Allogeneic hematopoietic stem cell transplantation

ARFCES:

Carcinoembryonic antigen

ASCT:

Allogeneic stem cell transplantation

BALF:

Bronchoalveolar lavage fluid

calcein AM:

Acetometoxy derivate of calcein

CCR5:

C-C chemokine receptor type 5

CXCL12:

Chemokine (C-X-C motif) ligand 12

CXCR-4:

C-X-C chemokine receptor type 4

DC:

Dendritic cell

EDTA:

Ethylenediaminetetraacetic acid

EM:

Electronic microscopy

EGFR:

Epidermal growth factor receptor

EMPs:

Endothelial microparticles

EpCAM:

Epithelial cell adhesion molecule

Er-MPs:

Erythrocyte-derived microparticles

ERK:

Extracellular signal-regulated kinase

Fas:

CD95

FasL:

Fas ligand

FMD:

Flow-mediated vasodilatation

GVHD:

Graft-versus-host disease

HER2:

Human epidermal growth receptor 2

HIV:

Human immunodeficiency virus

HSP:

Heat shock protein

HUVEC:

Human umbilical vein endothelial cell

ICAM-1:

Intercellular adhesion molecule 1

LLC:

Lewis lung carcinoma

LPS:

Lipopolysaccharide

MDR:

Multiple drug resistance

mRNA:

Messenger RNA

miRNA:

microRNA

MMPs:

Metalloproteinase

MPs:

Microparticles

NTA:

Nanoparticle tracking assay

PAF:

Platelet-activating factor

PBMCs:

Peripheral blood mononuclear cells

PCOS:

Polycystic ovary syndrome

P-gp:

P-glycoprotein

PMPs:

Platelet-derived microparticles

PMNs:

Polympophnonuclear neutrophils

PS:

Phosphatidylserine

RNA:

Ribonucleic acid

sPLA2:

Secretory phospholipase A2

PTEN:

Phosphatase and tensin homolog

ROCK-1:

Kinase

Rho-1:

Associated kinase

SIRS:

Systemic inflammatory response syndrome

SLE:

Systemic lupus erythematosus

STAT:

Signal transducer and activator of transcription

TF:

Tissue factor

TLR:

Toll-like receptor

TNF-α:

Tumor-necrotic factor alpha

TRAIL:

TNF-related apoptosis-induced ligand

TRM:

Transplantation-related mortality

TSG101:

Tumor specific antigen 101

VEGF:

Vascular endothelial growth factor.

References

  1. Corrado C, Raimondo S, Chiesi A, Ciccia F, De Leo G, Alessandro R: Exosomes as intercellular signaling organelles involved in health and disease: basic science and clinical applications. Int J Mol Sci. 2013, 14: 5338-5366. 10.3390/ijms14035338.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  2. Mathivanan S, Ji H, Simpson RJ: Exosomes: extracellular organelles important in intracellular communication. J Proteomics. 2010, 73: 1907-1920. 10.1016/j.jprot.2010.06.006.

    Article  CAS  PubMed  Google Scholar 

  3. Bobrie A, Colombo M, Raposo G, Thery C: Exosome secretion: molecular mechanisms and roles in immune responses. Traffic. 2011, 12: 1659-1668. 10.1111/j.1600-0854.2011.01225.x.

    Article  CAS  PubMed  Google Scholar 

  4. Chaput N, Thery C: Exosomes: immune properties and potential clinical implementations. Semin Immunopathol. 2011, 33: 419-440. 10.1007/s00281-010-0233-9.

    Article  CAS  PubMed  Google Scholar 

  5. Silverman JM, Reiner NE: Exosomes and other microvesicles in infection biology: organelles with unanticipated phenotypes. Cell Microbiol. 2011, 13: 1-9. 10.1111/j.1462-5822.2010.01537.x.

    Article  CAS  PubMed  Google Scholar 

  6. Lai CP, Breakefield XO: Role of exosomes/microvesicles in the nervous system and use in emerging therapies. Front Physiol. 2012, 3: 228-

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  7. Mause SF, Weber C: Microparticles: protagonists of a novel communication network for intracellular information exchange. Circ Res. 2010, 107: 1047-1057. 10.1161/CIRCRESAHA.110.226456.

    Article  CAS  PubMed  Google Scholar 

  8. van der Pol E, Boeing AN, Harrison P, Sturk A, Nieuwland R: Classification, functions, and clinical relevance of extracellular vesicles. Pharmacol Rev. 2012, 64: 676-705. 10.1124/pr.112.005983.

    Article  CAS  PubMed  Google Scholar 

  9. Torrecilhas AC, Schimacher RI, Alves MJM, Colli W: Vesicles as carriers of virulence factors in parasitic protozoan diseases. Microb Infect. 2012, 14: 1465-1474. 10.1016/j.micinf.2012.07.008.

    Article  CAS  Google Scholar 

  10. Flaumenhaft R: Formation and fate of platelet microparticles. Blood Cells Mol Dis. 2006, 36: 182-187. 10.1016/j.bcmd.2005.12.019.

    Article  CAS  PubMed  Google Scholar 

  11. Distler JH, Huber LC, Gay S, Distler O, Pisetsky DS: Microparticles as regulators of inflammation: novel players of cellular crosstalk in the rheumatic diseases. Arthritis Rheum. 2005, 52: 3337-3348. 10.1002/art.21350.

    Article  CAS  PubMed  Google Scholar 

  12. Shet AS: Characterizing blood microparticles: technical aspects and challenges. Vasc Health Risk Manag. 2008, 4: 769-774.

    PubMed Central  PubMed  Google Scholar 

  13. Beyer C, Pisetsky DS: The role of microparticles in the pathogenesis of rheumatic diseases. Nat Rev Rheumatol. 2010, 6: 21-29. 10.1038/nrrheum.2009.229.

    Article  CAS  PubMed  Google Scholar 

  14. Morel O, Morel N, Jesel L, Freyssinet JM, Toti F: Microparticles: a critical component in the nexus between inflammation, immunity and thrombosis. Semin Immunopathol. 2011, 33: 469-486. 10.1007/s00281-010-0239-3.

    Article  CAS  PubMed  Google Scholar 

  15. Zahra S, Anderson JA, Stirling D, Ludlam CA: Microparticles, malignancy and thrombosis. Br J Haematol. 2011, 152: 688-700. 10.1111/j.1365-2141.2010.08452.x.

    Article  CAS  PubMed  Google Scholar 

  16. Chaput N, Taieb J, Schartz NE, Andre F, Angevin E, Zitvogel L: Exosome-based immunotherapy. Cancer Immunol Immunother. 2004, 53: 234-239. 10.1007/s00262-003-0472-x.

    Article  CAS  PubMed  Google Scholar 

  17. Zwaal RF, Schroit AJ: Pathophysiologic implications of membrane phospholipid asymmetry in blood cells. Blood. 1997, 89: 1121-1132.

    CAS  PubMed  Google Scholar 

  18. Heijnen HF, Schiel AE, Fijnheer R, Geuze HJ, Sixma JJ: Activated platelets release two types of membrane vesicles: microvesicles by surface shedding and exosomes derived from exocytosis of multivesicular bodies and alpha-granules. Blood. 1999, 94: 3791-3799.

    CAS  PubMed  Google Scholar 

  19. Thery C, Boussac M, Veron P, Ricciardi-Castagnoli P, Raposo G, Garin J, Amigorena S: Proteomic analysis of denritic cell-derived exosomes: a secreted subcellular compartment distinct from apoptotic vesicles. J Immunol. 2001, 166: 7309-7318.

    Article  CAS  PubMed  Google Scholar 

  20. Orozco AF, Lewis DE: Flow cytometric analysis of circulating microparticles in plasma. Cytometry A. 2010, 77: 502-514.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  21. Yuana Y, Oosterkamp TH, Bahatyrova S, Ashcroft B, Garcia Rodriguez P, Bertina RM, Osanto S: Atomic force microscopy: a novel approach to the detection of nanosized blood microparticles. J Thromb Haemost. 2010, 8: 315-323. 10.1111/j.1538-7836.2009.03654.x.

    Article  CAS  PubMed  Google Scholar 

  22. Thery C, Zitvogel L, Amigorena S: Exosomes: composition, biogenesis and function. Nat Rev Immunol. 2002, 2: 569-579.

    CAS  PubMed  Google Scholar 

  23. Jy W, Horstmann LL, Jimenez JJ, Ahn JS, Biro E, Nieuwland R, Sturk A, Dignat-George F, Sabatier F, Camoin-Jau L, Sampol J, Hugel B, Zobairi F, Freyssinet JM, Nomura S, Shet AS, Key NS, Hebbel RP: Measuring circulating cell-derived microparticles. J Thromb Haemost. 2004, 2: 1842-1851. 10.1111/j.1538-7836.2004.00936.x.

    Article  CAS  PubMed  Google Scholar 

  24. Shah MD, Bergeron AL, Dong JF, Lopez JA: Flow cytometric measurement of microparticles: pitfalls and protocol modifications. Platelets. 2008, 19: 365-372. 10.1080/09537100802054107.

    Article  CAS  PubMed  Google Scholar 

  25. Dignat-George F, Freyssinet JM, Key NS: Centrifugation is a crucial step impacting microparticle measurement. Platelets. 2009, 20: 225-226. 10.1080/09537100902795500.

    Article  CAS  PubMed  Google Scholar 

  26. Dey-Hazra E, Hertel B, Kirsch T, Woywodt A, Lovric S, Haller H, Haubitz M, Erdbruegger U: Detection of circulating microparticles by flow cytometry: influence of centrifugation, filtration of buffer, and freezing. Vasc Health Risk Manag. 2010, 6: 1125-1133.

    PubMed Central  PubMed  Google Scholar 

  27. Gyorgy B, Modos K, Pallinger E, Paloczi K, Pasztoi M, Misjak P, Deli MA, Sipos A, Szalai A, Voszka I, Polgar A, Toth K, Csete M, Nagy G, Gay S, Falus A, Kittel A: Detection and isolation of cell-derived microparticles are compromised by protein complexes resulting from shared biophysical parameters. Blood. 2011, 117: e39-e48. 10.1182/blood-2010-09-307595.

    Article  CAS  PubMed  Google Scholar 

  28. Kirschner MB, Kao SC, Edelman JJ, Armstrong NJ, Vallely MP, Van Zandwijk N, Reid G: Haemolysis during sample preparation alters microRNA content of plasma. PLoS One. 2011, 6: e24145-10.1371/journal.pone.0024145.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  29. Lawrie AS, Albanyan A, Cardigan RA, Mackie IJ, Harrison P: Microparticle sizing by dynamic light scattering in fresh-frozen plasma. Vox Sang. 2009, 96: 206-212. 10.1111/j.1423-0410.2008.01151.x.

    Article  CAS  PubMed  Google Scholar 

  30. Van Ierssel SH, Van Craenenbroeck EM, Conraads VM, Van Tendeloo VF, Vrints CJ, Jorens PG, Hoymans VY: Flow cytometric detection of endothelial microparticles (EMP): effects of centrifugation and storage alter with the phenotype studied. Thromb Res. 2010, 125: 332-339. 10.1016/j.thromres.2009.12.019.

    Article  CAS  PubMed  Google Scholar 

  31. Van Der Pol E, Hoekstra AG, Sturk A, Otto C, Van Leeuwen TG, Nieuwland R: Optical and non-optical methods for detection and characterization of microparticles and exosomes. J Thromb Haemost. 2010, 8: 2596-2607. 10.1111/j.1538-7836.2010.04074.x.

    Article  CAS  PubMed  Google Scholar 

  32. Hess C, Sadallah S, Hefti A, Landmann R, Schifferli JA: Ectosomes released by human neutrophils are specialized functional units. J Immunol. 1999, 163: 4564-4573.

    CAS  PubMed  Google Scholar 

  33. Tilley RE, Holscher T, Belani R, Nieva J, Mackman N: Tissue factor activity is increased in a combined platelet and microparticle sample from cancer patients. Thromb Res. 2008, 122: 604-609. 10.1016/j.thromres.2007.12.023.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  34. Rood IM, Deegens JK, Merchant ML, Tamboer WP, Wilkey DW, Wetzels JF, Klein JB: Comparison of three methods for isolation of urinary microvesicles to identify biomarkers of nephrotic syndrome. Kidney Int. 2010, 78: 810-816. 10.1038/ki.2010.262.

    Article  CAS  PubMed  Google Scholar 

  35. Duarte TA, Noronha-Dutra AA, Nery JS, Ribeiro SB, Pitanga TN, Lapa e Silva JR, Arruda S, Boechat N: Mycobacterium tuberculosis-induced neutrophil ectosomes decrease macrophage activation. Tuberculosis (Edinb). 2012, 92: 218-225. 10.1016/j.tube.2012.02.007.

    Article  CAS  Google Scholar 

  36. Distler JH, Juengel A, Huber LC, Seemayer CA, Reich CF, Gay RE, Michel BA, Fontana A, Gay S, Pisetsky DS, Distler O: The induction of matrix metalloproteinase and cytokine expression in synovial fibroblasts stimulated with immune cell microparticles. Proc Natl Acad Sci USA. 2005, 102: 2892-2897. 10.1073/pnas.0409781102.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  37. Lima LG, Chammas R, Monteiro RQ, Moreira ME, Barcinski MA: Tumor-derived microvesicles modulate the establishment of metastatic melanoma in a phosphatidylserine-dependent manner. Cancer Lett. 2009, 283: 168-175. 10.1016/j.canlet.2009.03.041.

    Article  CAS  PubMed  Google Scholar 

  38. Witek RP, Yang L, Liu R, Jung Y, Omenetti A, Syn WK, Choi SS, Cheong Y, Fearing CM, Agboola KM, Chen W, Diehl AM: Liver cell-derived microparticles activate hedgehog signaling and alter gene expression in hepatic endothelial cells. Gastroenterology. 2009, 136: 320-330. 10.1053/j.gastro.2008.09.066.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  39. Porro C, Lepore S, Trotta T, Castellani S, Ratclif L, Battaglino A, Di Gioia S, Martinez MC, Conese M, Maffione AB: Isolation and characterization of microparticles in sputum from cystic fibrosis patients. Respir Res. 2010, 11: 94-10.1186/1465-9921-11-94.

    Article  PubMed Central  PubMed  Google Scholar 

  40. Philippova M, Suter Y, Toggweiler S, Schoenenberger AW, Joshi MB, Kyriakakis E, Erne P, Resink TJ: T-cadherin is present on endothelial microparticles and is elevated in plasma in early atherosclerosis. Eur Heart J. 2011, 32: 760-771. 10.1093/eurheartj/ehq206.

    Article  CAS  PubMed  Google Scholar 

  41. Mrvar-Brecko A, Sustar V, Jansa V, Stukelj R, Jansa R, Mujagic E, Kruljc P, Iglic A, Haegerstrand H, Kralj-Iglic V: Isolated microparticles from peripheral blood and body fluids as observed by scanning electron microscope. Blood Cells Mol Dis. 2010, 44: 307-312. 10.1016/j.bcmd.2010.02.003.

    Article  CAS  PubMed  Google Scholar 

  42. Burger D, Montezano AC, Nishigaki N, He Y, Carter A, Touyz RM: Endothelial microparticle formation by angiotensin II is via Ang II Receptor type I/NADPH oxidase/Rho kinase pathways targeted to lipid rafts. Arterioscler Thromb Vasc Biol. 2011, 31: 1898-1907. 10.1161/ATVBAHA.110.222703.

    Article  CAS  PubMed  Google Scholar 

  43. Miguet L, Sanglier S, Schaeffer C, Potier N, Mauvieux L, Van Dorsselaer A: Microparticles: a new tool for plasma membrane sub-cellular proteomic. Subcell Biochem. 2007, 43: 21-34. 10.1007/978-1-4020-5943-8_3.

    Article  PubMed  Google Scholar 

  44. Smalley DM, Ley K: Plasma-derived microparticles for biomarker discovery. Clin Lab. 2008, 54: 67-79.

    CAS  PubMed  Google Scholar 

  45. Street JM, Barran PE, Mackay CL, Weidt S, Balmforth C, Walsh TS, Chalmers RT, Webb DJ, Dear JW: Identification and proteomic profiling of exosomes in human cerebrospinal fluid. J Transl Med. 2012, 10: 5-10.1186/1479-5876-10-5.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  46. van der Pol E, Van Gemert MJ, Sturk A, Nieuwland R, Van Leeuwen TG: Single vs swarm detection of microparticles and exosomes by flow cytometry. J Thromb Haemost. 2012, 10: 919-930. 10.1111/j.1538-7836.2012.04683.x.

    Article  CAS  PubMed  Google Scholar 

  47. Jy W, Horstman KK, Ahn YS: Microparticle size and its relation to composition, functional activity, and clinical significance. Semin Thromb Hemost. 2010, 36: 876-880. 10.1055/s-0030-1267041.

    Article  PubMed  Google Scholar 

  48. Rubin O, Crettaz D, Tissot JD, Lion N: Pre-analytical and methodological challenges in red blood cell microparticle proteomics. Talanta. 2010, 82: 1-8. 10.1016/j.talanta.2010.04.025.

    Article  CAS  PubMed  Google Scholar 

  49. Xu Y, Nakane N, Maurer-Spurej E: Novel test for microparticles in platelet-rich plasma and platelet concentrates using dynamic light scattering. Transfusion. 2011, 51: 363-370. 10.1111/j.1537-2995.2010.02819.x.

    Article  PubMed  Google Scholar 

  50. Tesselaar ME, Romijin FP, Van Der Linden IK, Prins FA, Bertina RM, Osanto S: Microparticle-associated tissue factor activity: a link between cancer and thrombosis?. J Thromb Haemost. 2007, 5: 520-527. 10.1111/j.1538-7836.2007.02369.x.

    Article  CAS  PubMed  Google Scholar 

  51. Manly DA, Wang J, Glover SL, Kasthuri R, Liebman HA, Key NS, Mackman N: Increased microparticle tissue factor activity in cancer patients with venous thromboembolism. Thromb Res. 2010, 125: 511-512. 10.1016/j.thromres.2009.09.019.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  52. Zwicker JI: Impedance-based flow cytometry for the measurement of microparticles. Semin Thromb Hemost. 2010, 36: 819-823. 10.1055/s-0030-1267035.

    Article  PubMed  Google Scholar 

  53. Zwicker JI, Lacroix R, Dignat-George F, Furie BC, Furie B: Measurement of platelet microparticles. Methods Mol Biol. 2012, 788: 127-139. 10.1007/978-1-61779-307-3_10.

    Article  CAS  PubMed  Google Scholar 

  54. Bernimoulin M, Waters EK, Foy M, Steele BM, Sullivan M, Falet H, Walsh MT, Barteneva N, Geng JG, Hartwig JH, Maguire PB, Wagner DD: Differential stimulation of monocytic cells results in distinct populations of microparticles. J Thromb Haemost. 2009, 7: 1019-1028. 10.1111/j.1538-7836.2009.03434.x.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  55. Cerri C, Chimenti D, Conti I, Neri T, Paggiaro P, Celi A: Monocyte/macrophage-derived microparticles up-regulate inflammatory mediator synthesis by human airway epithelial cells. J Immunol. 2006, 177: 1975-1980.

    Article  CAS  PubMed  Google Scholar 

  56. Enjeti AK, Lincz L, Seldon M: Bio-maleimide as a generic stain for detection and quantitation of microparticles. Int J Lab Hematol. 2008, 30: 196-199. 10.1111/j.1751-553X.2007.00937.x.

    Article  CAS  PubMed  Google Scholar 

  57. Graves LE, Ariztia EV, Navari JR, Matzel HJ, Stack MS, Fishman DA: Proinvasive properties of ovarian cancer ascites-derived membrane vesicles. Cancer Res. 2004, 64: 7045-7049. 10.1158/0008-5472.CAN-04-1800.

    Article  CAS  PubMed  Google Scholar 

  58. Piccin A, Murphy WG, Smith OP: Circulating microparticles: pathophysiology and clinical implications. Blood Rev. 2007, 21: 157-171. 10.1016/j.blre.2006.09.001.

    Article  CAS  PubMed  Google Scholar 

  59. Smalley DM, Sheman NE, Nelson K, Theodorescu D: Isolation and identification of potential urinary microparticle biomarkers of bladder cancer. J Proteome Res. 2008, 7: 2088-2096. 10.1021/pr700775x.

    Article  CAS  PubMed  Google Scholar 

  60. Lescuyer P, Pernin A, Hainard A, Bigeire C, Burgess JA, Zimmerman-Ivol C, Sanchez JC, Schifferli JA, Hochstrasser DF, Moll S: Proteomics analysis of a podocyte vesicle-enriched fraction from normal human and pathological urines. Proteomics Clin Appl. 2008, 2: 1008-1018. 10.1002/prca.200800033.

    Article  CAS  PubMed  Google Scholar 

  61. Berckmans RJ, Sturk A, Van Tienen LM, Schaap MC, Nieuwland R: Cell-derived vesicles exposing coagulant tissue factor in saliva. Blood. 2011, 117: 3172-3180. 10.1182/blood-2010-06-290460.

    Article  CAS  PubMed  Google Scholar 

  62. Uszynski M, Zekanowska E, Uszynski W, Kuczynski J, Zylinski A: Microparticles (MPs), tissue factor (TF) and tissue factor inhibitor (TFPI) in cord blood plasma. A preliminary study and literature survey of procoagulant properties of MPs. Eur J Obstet Gynecol Reprod Biol. 2011, 158: 37-41. 10.1016/j.ejogrb.2011.04.026.

    Article  CAS  PubMed  Google Scholar 

  63. Bastarche JA, Fremont RD, Kropski JA, Bossert FR, Ware LB: Procoagulant alveolar microparticles in the lungs of patients with acute respiratory distress syndrome. Am J Physiol Lung Cell Mol Physiol. 2009, 297: L1035-L1041. 10.1152/ajplung.00214.2009.

    Article  CAS  Google Scholar 

  64. Guervilly C, Lacroix R, Forel JM, Roch A, Camoin-Jau L, Papazian L, Dignat-George F: High levels of circulating leukocyte microparticles are associated with better outcome in acute respiratory distress syndrome. Crit Care. 2011, 15: R31-10.1186/cc9978.

    Article  PubMed Central  PubMed  Google Scholar 

  65. Kockx MM: Apoptosis in the atherosclerotic plaque: quantitative and qualitative aspects. Arterioscler Thromb Vasc Biol. 1998, 18: 1519-1522. 10.1161/01.ATV.18.10.1519.

    Article  CAS  PubMed  Google Scholar 

  66. Mallat Z, Hugel B, Ohan J, Leseche G, Freyssinet JM, Tedgui A: Shed membrane microparticles with procoagulant potential in human atherosclerotic plaques: a role of apoptosis in plaque thrombogenecity. Circulation. 1999, 99: 348-353. 10.1161/01.CIR.99.3.348.

    Article  CAS  PubMed  Google Scholar 

  67. Leroyer AS, Tedgui A, Boulanger CM: Microparticles and type 2 diabetes. Diab Metab. 2008, 34: S27-S31.

    Article  CAS  Google Scholar 

  68. Sadallah S, Lach E, Lutz HU, Schwarz S, Guerne PA, Schifferli JA: CR1, CD35 in synovial fluid from patients with inflammatory joint diseases. Arthritis Rheum. 1997, 40: 520-526. 10.1002/art.1780400318.

    Article  CAS  PubMed  Google Scholar 

  69. Berckmans RJ, Nieuwland R, Kraan MC, Schaap MC, Pots D, Smeets TJ, Sturk A, Tak PP: Synovial microparticles from arthritic patients modulate chemokine and cytokine release by synoviocytes. Arthritis Res Ther. 2005, 7: R536-R544. 10.1186/ar1706.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  70. Biro E, Nieuwland R, Tak PP, Pronk LM, Schaap MC, Sturk A, Hack CE: Activated complement compounds and complement activator molecules on the surface of cell-derived microparticles in patients with rheumatoid arthritis and healthy individuals. Ann Rheum Dis. 2007, 66: 1085-1092. 10.1136/ard.2006.061309.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  71. Messer L, Alsaleh G, Freyssinet JM, Zobairi F, Leray I, Gottenberg JE, Sibilia J, Toti-Orfanoudakis F, Wachsmann D: Microparticle-induced release of B-lymphocyte regulators by rheumatoid synoviocytes. Arthritis Res Ther. 2009, 11: R40-10.1186/ar2648.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  72. Chahed S, Leroyer AS, Benzerroug M, Gaucher D, Georguescu A, Picaud S, Silvestre JS, Gaudric A, Tedgui A, Massin P, Boulanger CM: Increased vitreous shedding of microparticles in proliferative diabetic retinopathy stimulates endothelial proliferation. Diabetes. 2010, 59: 694-701. 10.2337/db08-1524.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  73. Aleman MM, Gardiner C, Harrison P, Wolberg AS: Differential contributions of monocyte- and platelet-derived microparticles towards thrombin generation and fibrin formation and stability. J Thromb Haemost. 2011, 9: 2251-2261. 10.1111/j.1538-7836.2011.04488.x.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  74. Berckmans RJ, Neiuwland R, Boeing AN, Romijn FP, Hack CE, Stark A: Cell-derived microparticles circulate in healthy humans and support low grade thrombin generation. Thromb Haemost. 2001, 85: 639-646.

    CAS  PubMed  Google Scholar 

  75. Flaumenhaft R, Dilks JR, Richardson J, Alden E, Patel-Hett SR, Battinelli E, Klement GL, Sola-Visner M, Italiano JE: Megakaryocyte-derived microparticles: direct visualization and distinction from platelet-derived microparticles. Blood. 2009, 113: 1112-1121.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  76. Siljander PR: Platelet-derived microparticles-an updated perspective. Thromb Res. 2011, 127: S30-S33.

    Article  CAS  PubMed  Google Scholar 

  77. Rank A, Nieuwland R, Delker R, Koehler A, Toth B, Pihusch V, Wilkowski R, Pihusch R: Cellular origin of platelet-derived microparticles in vivo. Thrombosis Res. 2010, 126: e255-e259. 10.1016/j.thromres.2010.07.012.

    Article  CAS  Google Scholar 

  78. Angelillo-Scherrer A: Leukocyte-derived microparticles in vascular homeostasis. Circ Res. 2012, 110: 356-369. 10.1161/CIRCRESAHA.110.233403.

    Article  CAS  PubMed  Google Scholar 

  79. Sadallah S, Eken C, Schifferli JA: Ectosomes as modulators of inflammation and immunity. Clin Exp Immunol. 2011, 163: 26-32. 10.1111/j.1365-2249.2010.04271.x.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  80. Sims PJ, Faioni EM, Wiedmer T, Shattil SJ: 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.

    CAS  PubMed  Google Scholar 

  81. Stein JMK, Luzio JP: Ectocytosis caused by sublytic autologous complement attack on human neutrophils. The sorting of endogenous plasma-membrane proteins and lipids into shed vesicles. Biochem J. 1991, 274: 381-386.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  82. Dolo V, Ginestra A, Cassara D, Violini S, Lucania G, Torrisi MR, Nagase H, Canevaari S, Pavan A, Vittorelli ML: Selective localization of matrix metalloproteinase 9, beta1 integrins, and human lymphocyte antigen class I molecules on membrane vesicles shed by 8701-BC breast carcinoma cells. Cancer Res. 1998, 58: 4468-4474.

    CAS  PubMed  Google Scholar 

  83. Biro E, Akkerman JW, Hoek FJ, Gorter G, Pronk LM, Sturk A, Nieuwland R: The phospholipid composition and cholesterol content of platelet-derived microparticles: a comparison with platelet membrane fractions. J Thromb Haemost. 2005, 3: 2754-2763. 10.1111/j.1538-7836.2005.01646.x.

    Article  CAS  PubMed  Google Scholar 

  84. Peterson DB, Sander T, Kaul S, Wakim BT, Halligan B, Twigger S, Pritchard KA, Oldham KT, Ou JS: Comparative proteomic analysis of PAI-1 and TNF-alpha-derived endothelial microparticles. Proteomics. 2008, 8: 2430-2446. 10.1002/pmic.200701029.

    Article  CAS  PubMed  Google Scholar 

  85. Jimenez JJ, Jy W, Mauro LM, Soderland C, Horstman LL, Ahn YS: Endothelial cells release phenotypically and quantitatively distinct microparticles in activation and apoptosis. Thromb Res. 2003, 109: 175-180. 10.1016/S0049-3848(03)00064-1.

    Article  CAS  PubMed  Google Scholar 

  86. Shai E, Rosa I, Parguina AF, Motahedeh S, Varon D, Garcia A: Comparative analysis of platelet-derived microparticles reveals differences in their amount and proteome depending on the platelet stimulus. J Proteomics. 2012, 76: 287-296.

    Article  CAS  PubMed  Google Scholar 

  87. Sinauridze EI, Kireev DA, Popenko NY, Pichugin AV, Panteleev MA, Krymskaya OV, Ataullakhanov FI: Platelet microparticle membranes have 50- to 100-fold higher specific procoagulant activity than activated platelets. Thromb Haemost. 2007, 97: 425-434.

    CAS  PubMed  Google Scholar 

  88. Pluskota E, Woody NM, Szpak D, Ballantyne CM, Soloviev DA, Simon DI, Plow EF: Expression, activation, and function of integrin alphaM/beta2 (Mac-1) on neutrophil-derived microparticles. Blood. 2008, 112: 2327-2335. 10.1182/blood-2007-12-127183.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  89. Del Conde I, Shrimpton CN, Thiagarajan P, Lopez JA: Tissue factor-bearing microvesicles arise from lipid rafts and fuse with activated platelets to initiate coagulation. Blood. 2005, 106: 1604-1611. 10.1182/blood-2004-03-1095.

    Article  CAS  PubMed  Google Scholar 

  90. Al-Nedawi K, Meehan K, Micallef J, Lhotak V, May L, Guha A, Rak J: Intercellular transfer of the oncogenic receptor EGFRvIII by microvesicles derived from tumour cells. Nat Cell Biol. 2008, 10: 619-624. 10.1038/ncb1725.

    Article  CAS  PubMed  Google Scholar 

  91. Daleke DL: Regulation of transbilayer plasma membrane phospholipid asymmetry. J Lipid Res. 2003, 44: 233-242. 10.1194/jlr.R200019-JLR200.

    Article  CAS  PubMed  Google Scholar 

  92. Bevers EM, Comfurious P, Dekkers DW, Harmsma M, Zwaal RF: Transmembrane phospholipid distribution in blood cells: control mechanisms and pathophysiological significance. Biol Chem. 1998, 379: 973-986.

    CAS  PubMed  Google Scholar 

  93. Perez-Pujol S, Marker PH, Key NS: Platelet microparticles are heterogeneous and highly dependent on the activation mechanism: studies using a new digital flow cytometer. Cytometry A. 2007, 71: 38-45.

    Article  PubMed  Google Scholar 

  94. Rukoyatkina N, Begonja AJ, Geiger J, Eigenthaler M, Walter U, Gambarayan S: Phosphatidylserine surface expression and integrin alpha IIb beta 3 activity on thrombin/convulxin stimulated platelets/particles of different sizes. Br J Haematol. 2009, 144: 591-602. 10.1111/j.1365-2141.2008.07506.x.

    Article  CAS  PubMed  Google Scholar 

  95. Key NS: Analysis of tissue factor positive microparticles. Thromb Res. 2010, 125: S42-S45.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  96. Siljander P, Farndale RW, Feijge MA, Comfurius P, Kos S, Bevers EM, Heemskerk JW: Platelet adhesion enhances the glycoprotein VI-dependent procoagulant response: involvement of p38 MAP kinase and calpain. Arterioscler Thromb Vasc Biol. 2001, 21: 618-627. 10.1161/01.ATV.21.4.618.

    Article  CAS  PubMed  Google Scholar 

  97. Shet AS, Aras O, Gupta K, Hass MJ, Rausch DJ, Saba N, Koopmeiners L, Key NS, Hebbel RP: Sickle blood contains tissue factor positive microparticles derived from endothelial cells and monocytes. Blood. 2003, 102: 2678-2683. 10.1182/blood-2003-03-0693.

    Article  CAS  PubMed  Google Scholar 

  98. Amabile N, Guerin AP, Leroyer A, Mallat Z, Nguyen C, Boddaert J, London GM, Tedgui A, Boulanger CM: Circulating microparticles are associated with vascular dysfunction in patients with end-stage renal failure. J Am Soc Nephrol. 2005, 16: 3381-3388. 10.1681/ASN.2005050535.

    Article  CAS  PubMed  Google Scholar 

  99. Fox JE, Austin CD, Reynolds CC, Steffen PK: Evidence that agonist-activated activation of calpain causes the shedding of procoagulant-containing microvesicles from the membrane of aggregating platelets. J Biol Chem. 1991, 266: 13289-13295.

    CAS  PubMed  Google Scholar 

  100. Muralidharan-Chari V, Hoover H, Clancy J, Schweitzer J, Suckow MA, Schroeder V, Castellino FJ, Schorey JS, D’Souza-Schorey C: ADP-ribosylation factor 6 regulates tumorigenic and invasive properties in vivo. Cancer Res. 2009, 69: 2201-2209. 10.1158/0008-5472.CAN-08-1301.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  101. Martinez MC, Martin C, Toti F, Fressinaud E, Dachary-Prigent J, Meyer D, Freyssinet JM: Significance of capacitative Ca2+ entry in the regulation of phosphatydylcholine expression at the surface of stimulated cells. Biochemistry. 1999, 38: 10092-10098. 10.1021/bi990129p.

    Article  CAS  PubMed  Google Scholar 

  102. Morel O, Jesel L, Freyssinet JM, Toti F: Cellular mechanisms underlying the formation of circulating microparticles. Arterioscler Thromb Vasc Biol. 2011, 31: 15-26. 10.1161/ATVBAHA.109.200956.

    Article  CAS  PubMed  Google Scholar 

  103. Lentz BR: Exposure of platelet membrane phosphatidylserine regulates blood coagulation. Prog Lipid Res. 2003, 42: 423-438. 10.1016/S0163-7827(03)00025-0.

    Article  CAS  PubMed  Google Scholar 

  104. Martinez MC, Tual-Chalot S, Leonetti D, Andriantsitohaina R: Microparticles: targets and tools in cardiovascular disease. Trends Pharmacol Sci. 2011, 32: 659-665. 10.1016/j.tips.2011.06.005.

    Article  CAS  PubMed  Google Scholar 

  105. Connor DE, Exner T, Ma DD, Joseph JE: The majority of circulating platelet-derived microparticles fail to bind annexin V, lack phospholipid-procoagulant activity and demonstrate greater expression of glycoprotein Ib. Thromb Haemost. 2010, 103: 1044-1052. 10.1160/TH09-09-0644.

    Article  CAS  PubMed  Google Scholar 

  106. Kim SJ, Moon GJ, Cho YH, Kang HY, Hyung NK, Kim D, Lee JH, Nam JY, Bang OY: Circulating mesenchymal stem cells microparticles in patients with cerebrovascular disease. PLoS One. 2012, 7: e37036-10.1371/journal.pone.0037036.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  107. Nielsen CT: Circulating microparticles in systemic lupus erythematosus. Dan Med J. 2012, 59: B4548-

    PubMed  Google Scholar 

  108. Hanayama R, Tanaka M, Miwa K, Shinohara A, Iwamatsu A, Nagata S: Identification of a factor that links apoptotic cells to phagocytes. Nature. 2002, 417: 182-187. 10.1038/417182a.

    Article  CAS  PubMed  Google Scholar 

  109. Miyanushi M, Tada K, Koike M, Uchiyama Y, Kitamura T, Nagata S: Identification of Tim4 as a phosphatidylserine receptor. Nature. 2007, 450: 435-439. 10.1038/nature06307.

    Article  CAS  Google Scholar 

  110. Freyssinet JM: Cellular microparticles: What are they bad or good for?. J Thromb Haemost. 2003, 1: 1655-1662. 10.1046/j.1538-7836.2003.00309.x.

    Article  CAS  PubMed  Google Scholar 

  111. VanWijk MJ, VanBavel E, Sturk A, Nieuwland R: Microparticles in cardiovascular diseases. Cardiovasc Res. 2003, 59: 277-287. 10.1016/S0008-6363(03)00367-5.

    Article  CAS  PubMed  Google Scholar 

  112. Popescu NI, Lupu C, Lupu F: Extracellular protein disulfide isomerase regulates coagulation on endothelial cells through modulation of phosphatidylserine exposure. Blood. 2010, 116: 993-1001. 10.1182/blood-2009-10-249607.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  113. Furlan-Freguia C, Marchese P, Gruber A, Ruggeri ZM, Ruf W: P2X7 receptor signaling contributes to tissue factor-dependent thrombosis in mice. J Clin Invest. 2011, 121: 2932-2944. 10.1172/JCI46129.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  114. Bianco F, Perrotta C, Novellino L, Francolini M, Riganti L, Menna E, Saglietti L, Schuchman EH, Furlan R, Clementi E, Matteoli M, Verderio C: Acid sphingomyelinase activity triggers microparticle release from glial cells. EMBO J. 2009, 28: 1043-1054. 10.1038/emboj.2009.45.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  115. Kunzelmann-Marche C, Freyssinet JM, Martinez MC: Loss of plasma membrane phospholipid asymmetry requires raft integrity. Role of transient receptor potential channels and ERK pathway. J Biol Chem. 2002, 277: 19876-19881. 10.1074/jbc.M200324200.

    Article  CAS  PubMed  Google Scholar 

  116. Lopez JA, Del Conde I, Shrimpton CN: Receptors, rafts, and microvesicles in thrombosis and inflammation. J Thromb Haemost. 2005, 3: 1737-1744. 10.1111/j.1538-7836.2005.01463.x.

    Article  CAS  PubMed  Google Scholar 

  117. Aoki N, Jin-No S, Nakagawa Y, Asai N, Arakawa E, Tamura N, Tamura T, Matsuda T: Identification and characterization of microvesicles secreted by 3T3-L1 adipocytes: redox- and hormone-dependent induction of milk fat globule-epidermal growth factor 8-associated microvesicles. Endocrinology. 2007, 148: 3850-3862. 10.1210/en.2006-1479.

    Article  CAS  PubMed  Google Scholar 

  118. Thery C, Ostrowski M, Segura E: Membrane vesicles as conveyors of immune responses. Nat Rev Immunol. 2009, 9: 581-593. 10.1038/nri2567.

    Article  CAS  PubMed  Google Scholar 

  119. Pisetsky DS, Spencer DM: Effects of progesterone and estradiol sex hormones on the release of microparticles by RAW 264.7 macrophages stimulated by Poly (I:C). Clin Vaccine Imm. 2011, 18: 1420-1426. 10.1128/CVI.05110-11.

    Article  CAS  Google Scholar 

  120. Tushuizen ME, Diamant M, Peypers EG, Hoek FJ, Heine RJ, Sturk A, Nieuwland R: Postprandial changes in phospholipid composition of circulating microparticles are not associated with coagulation activation. Thromb Res. 2001, 130: 115-121.

    Article  CAS  Google Scholar 

  121. Nomura S, Tandon NN, 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. 10.1016/S0021-9150(01)00433-6.

    Article  CAS  PubMed  Google Scholar 

  122. Jayachanadran M, Litwiller RD, Owen WG, Miller VM: Circulating microparticles and endogenous estrogen in newly menopausal women. Climacteric. 2009, 12: 177-184. 10.1080/13697130802488607.

    Article  CAS  Google Scholar 

  123. Shirafuji T, Hamaguchi H, Higuchi M, Kanda F: Measurement of platelet-derived microparticle levels using an enzyme-linked immunosorbent assay in polymyositis and dermatomyosistis patients. Muscle Nerve. 2009, 39: 586-590. 10.1002/mus.21311.

    Article  CAS  PubMed  Google Scholar 

  124. Gerrits AJ, Koekman CA, Yildirim C, Nieuwland R, Akkerman JW: Insulin inhibits tissue factor expression in monocytes. J Thromb Haemost. 2008, 7: 198-205.

    Article  PubMed  CAS  Google Scholar 

  125. Bergsmedh A, Szeles A, Henriksson M, Bratt A, Folkman MJ, Spetz AL, Holmgren L: Horizontal transfer of oncogenes by uptake of apoptotic bodies. Proc Natl Acad Sci USA. 2001, 98: 6407-6411. 10.1073/pnas.101129998.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  126. Valadi H, Ekstroem K, Bossios A, Sjoestrand M, Lee JJ, Loetvall JO: Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat Cell Biol. 2007, 9: 654-659. 10.1038/ncb1596.

    Article  CAS  PubMed  Google Scholar 

  127. Risitano A, Beaulieu LM, Vitseva O, Freedman JE: Platelets and platelet-like particles mediate intercellular RNA transfer. Blood. 2012, 119: 6288-6295. 10.1182/blood-2011-12-396440.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  128. Hunter MP, Ismail N, Zhang X, Aguda BD, Lee EJ, Yu L, Xiao T, Schafer J, Lee ML, Schmittgen TD, Nana-Sinkam SP, Jarjoura D, Marsh CB: Detection of microRNA expression in human peripheral blood microvesicles. PLoS One. 2008, 3: e3694-10.1371/journal.pone.0003694.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  129. Teruel R, Corral J, Perez-Andreu V, Martinez-Martinez I, Vicente V, Martinez C: Potential role of miRNAs in developmental haemostasis. PLoS One. 2011, 6: e17648-10.1371/journal.pone.0017648.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  130. Diehl P, Fricke A, Sander L, Stamm J, Bassler N, Htun N, Ziemann M, Helbing T, El-Osta A, Jowett JB, Peter K: Microparticles: major transport vehicles for distinct miRNAs in circulation. Cardiovasc Res. 2012, 93: 633-634. 10.1093/cvr/cvs007.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  131. Mueller G, Schneider M, Biemer-Daub G, Wied S: Microvesicles released from rat adipocytes and harboring glycophosphatidylinositol-anchored proteins transfer RNA stimulating lipid synthesis. Cell Signal. 2011, 23: 1207-1223. 10.1016/j.cellsig.2011.03.013.

    Article  CAS  Google Scholar 

  132. Cantaluppi V, Gatti S, Medica D, Figliolini F, Bruno S, Deregibus MC, Sordi A, Biancone L, Tetta C, Camussi G: Microvesicles derived from endothelial progenitor cells protect the kidney from ischemia-reperfusion injury by microRNA-dependent reprogramming of resident renal cells. Kidney Int. 2012, 82: 412-427. 10.1038/ki.2012.105.

    Article  CAS  PubMed  Google Scholar 

  133. Hergenreider E, Heydt S, Treguer K, Boettger T, Horrevoets AJG, Zeiher AM, Scheffer MP, Frangakis AS, Yin X, Mayr M, Braun T, Urbich C, Boon RA, Dimmeler S: Atheroprotective communication between endothelial cells and smooth muscle cells through miRNA. Nat Cell Biology. 2012, 14: 249-256. 10.1038/ncb2441.

    Article  CAS  PubMed  Google Scholar 

  134. Watanabe J, Marathe GK, Neilsen PO, Weyrich AS, Harrison KA, Murphy RC, Zimmerman GA, McInthyre TM: Endotoxins stimulate neuthrophil adhesion followed by synthesis and release of platelet-activating factor in microparticles. J Biol Chem. 2003, 278: 33161-33168. 10.1074/jbc.M305321200.

    Article  CAS  PubMed  Google Scholar 

  135. Cardo LJ, Wilder D, Salata J: Neuthrophil priming, caused by cell membranes and microvesicles in packed red blood cell units, is abrogated by leukocyte depletion at collection. Transfus Apher Sci. 2008, 38: 117-125. 10.1016/j.transci.2008.01.004.

    Article  PubMed  Google Scholar 

  136. Fujimi S, Ogura H, Tanaka H, Koh T, Hosotsubo H, Nakamori Y, Kuwagata Y, Shimazu T, Sugimoto H: Increased production of leukocyte microparticles with enhanced expression of adhesion molecules from activated polymorphnonuclear leukocytes in severely injured patients. J Trauma. 2003, 54: 114-119. 10.1097/00005373-200301000-00014.

    Article  PubMed  Google Scholar 

  137. Press JZ, Reyes M, Pitteri SJ, Pennil C, Garcia R, Goff BA, Hanash SM, Swisher EM: Microparticles from ovarian carcinomas are shed into ascites and promote cell migration. Int J Gynecol Cancer. 2012, 22: 546-552. 10.1097/IGC.0b013e318241d9b9.

    Article  PubMed  Google Scholar 

  138. Rautou PE, Leroyer AS, Ramkhelawon B, Devue C, Duflaut D, Vion AC, Nalbone G, Castier Y, Leseche G, Lehoux S, Tedgui A, Boulanger CM: Microparticles from human atherosclerotic plaques promote endothelial ICAM-1-dependent monocyte adhesion and transendothelial migration. Circ Res. 2011, 108: 335-343. 10.1161/CIRCRESAHA.110.237420.

    Article  CAS  PubMed  Google Scholar 

  139. Sadallah S, Eken C, Schifferli JA: Erythrocyte-derived ectosomes have immunosuppressive properties. J Leukoc Biol. 2008, 84: 1316-1325. 10.1189/jlb.0108013.

    Article  CAS  PubMed  Google Scholar 

  140. Gasser O, Schifferli JA: Microparticles released by human neutrophils adhere to erythrocytes in the presence of complement. Exp Cell Res. 2005, 307: 381-387. 10.1016/j.yexcr.2005.03.011.

    Article  CAS  PubMed  Google Scholar 

  141. Mack M, Kleinschmidt A, Bruehl H, Klier C, Nelson PJ, Cihak J, Plachy J, Stangassinger M, Erfle V, Schloendorff D: Transfer of the chemokine receptor CCR5 between cells by membrane-derived microparticles: a mechanism for cellular immunodeficiency virus 1 infection. Nat Med. 2000, 6: 769-775. 10.1038/77498.

    Article  CAS  PubMed  Google Scholar 

  142. Giesen PL, Rauch BA, Bohrmann B, Kling D, Rogue M, Fallon JT, Badimon JJ, Himber J, Riederer MA, Nemerson Y: Blood-borne tissue factor: another view of thrombosis. Proc Natl Acad Sci USA. 1999, 96: 2311-2315. 10.1073/pnas.96.5.2311.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  143. Huber V, Fais S, Iero M, Lugini L, Canese P, Squarcina P, Zaccheddu A, Colone M, Arancia G, Gentile M, Seregni E, Valenti R, Ballabio G, Belli F, Leo E, Parmiani G, Rivoltini L: Human colorectal cancer cells induce T-cell death through release of proapoptotic microvesicles: role in immune escape. Gastroenterology. 2005, 128: 1796-1804. 10.1053/j.gastro.2005.03.045.

    Article  CAS  PubMed  Google Scholar 

  144. Lee TH, D’Asti E, Magnus N, Al-Nedawi K, Meehan B, Rak J: Microvesicles as mediators of intercellular communication in cancer-the emerging science of cellular “debris”. Semin Immunopathol. 2011, 33: 455-467. 10.1007/s00281-011-0250-3.

    Article  PubMed  Google Scholar 

  145. Rozmyslowicz T, Majka M, Kijowski J, Murphy SL, Conover DO, Poncz M, Ratajczak J, Gaulton GN, Ratajczak MZ: Platelet and megakariocyte-derived microparticles transfer CXCR4 receptor to CXCR-null cells and make them susceptible to infection by X4-HIV. AIDS. 2003, 17: 33-42. 10.1097/00002030-200301030-00006.

    Article  CAS  PubMed  Google Scholar 

  146. Zernecke A, Bidzhekov K, Noels H, Shagdarsuren E, Gan L, Denecke B, Hristov M, Koeppel T, Jahantigh MN, Lutgens E, Wang S, Olson EN, Schober A, Weber C: Delivery of microRNA-126 by apoptotic bodies induces CXCl12-dependent vascular protection. Sci Signal. 2009, 2: ra81-10.1126/scisignal.2000610.

    Article  PubMed  Google Scholar 

  147. Distler JH, Akhmetshina A, Dees C, Jüngel A, Stürzl M, Gay S, Pisetsky DS, Schett G, Distler O: Induction of apoptosis in circulating angiogenic cells by microparticles. Arthritis Rheum. 2011, 63: 2067-2077. 10.1002/art.30361.

    Article  CAS  PubMed  Google Scholar 

  148. Aliotta JM, Pereira M, Johnson KW, De Paz N, Dooner MS, Puente N, Ayala C, Brilliant K, Berz D, Lee D, Ramratnam B, McMillan PN, Hixson DC, Josic D, Quesenberry PJ: Microvesicle entry into marrow cells mediates tissue-specific changes in mRNA by direct delivery of mRNA and induction of transcription. Exp Hematol. 2010, 38: 233-245. 10.1016/j.exphem.2010.01.002.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  149. Deregibus MC, Tetta C, Camussi G: The dynamic stem cell microenvironment is orchestrated by microvesicle-mediated transfer of genetic information. Histol Histopathol. 2010, 25: 397-404.

    CAS  PubMed  Google Scholar 

  150. Ratajczak J, Miekus K, Kucia M, Zhang J, Reca R, Dvorak P, Ratajczak MZ: Embryonic stem cell-derived microvesicles reprogram hematopoietic progenitors: evidence for horizontal transfer of mRNA and protein delivery. Leukemia. 2006, 20: 847-856. 10.1038/sj.leu.2404132.

    Article  CAS  PubMed  Google Scholar 

  151. Herrera MB, Fonsato V, Gatti S, Deregibus MC, Sordi A, Cantarella D, Calogero R, Bussolati B, Tatta C, Camussi G: Human liver stem cell-derived microvesicles accelerate hepatic regeneration in hepatectomized rats. J Cell Mol Med. 2010, 14: 1605-1618.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  152. Yuan A, Farber EL, Rapoport AL, Tejada D, Deniskin R, Akhmedov NB, Farber DB: Transfer of microRNAs by embryonic stem cell microvesicles. PLoS One. 2009, 4: e4722-10.1371/journal.pone.0004722.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  153. Collino F, Deregibus MC, Bruno S, Sterpone L, Aghemo G, Viltono L, Tetta C, Camussi G: Microvesicles derived from adult human bone marrow and tissue specific mesenchymal stem cells shuttle selected pattern of miRNAs. PLoS One. 2010, 5: e11803-10.1371/journal.pone.0011803.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  154. Sarkar A, Mitra S, Mehta S, Raices R, Wewers MD: Monocyte derived microvesicles deliver a cell death message via encapsulated caspase-1. PLoS One. 2009, 4: e7140-10.1371/journal.pone.0007140.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  155. Abid Hussein MN, Nieuwland R, Hau CM, Evers LM, Meesters EW, Sturk A: Cell-derived microparticles contain caspase 3 in vitro and in vivo. J Thromb Haemost. 2005, 3: 888-896. 10.1111/j.1538-7836.2005.01240.x.

    Article  CAS  PubMed  Google Scholar 

  156. Abid Hussein MN, Boeing AN, Sturk A, Hau CM, Nieuwland R: Inhibition of microparticle release triggers endothelial cell apoptosis and detachment. Thromb Haemost. 2007, 98: 1096-1107.

    PubMed  Google Scholar 

  157. Boeing AN, Hau CM, Sturk A, Nieuwland R: Platelet microparticles contain active caspase 3. Platelets. 2008, 19: 96-103. 10.1080/09537100701777295.

    Article  CAS  Google Scholar 

  158. Albanese J, Meterissian S, Kontogiannea M, Dubreuil C, Hand A, Sorba S, Dainiak N: Biologically active Fas antigen and its cognate ligand are expressed on plasma membrane–derived extracellular vesicles. Blood. 1998, 91: 3862-3874.

    CAS  PubMed  Google Scholar 

  159. Andreola G, Rivoltini L, Castelli C, Huber V, Perego P, Deho P, Sguarcina P, Accornero P, Lozupone F, Lugini L, Stringaro A, Molinari A, Arancia G, Gentile M, Parmiani G, Fais S: Induction of lymphocyte apoptosis by tumor cell secretion of FASL-bearing microvesicles. J Exp Med. 2002, 195: 1303-1316. 10.1084/jem.20011624.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  160. Taylor DD, Gercel-Taylor C: Tumour-derived exosomes and their role in cancer-associated T-cell signalling defects. Br J Cancer. 2005, 92: 305-311.

    PubMed Central  CAS  PubMed  Google Scholar 

  161. Kim JW, Wieckowski E, Taylor DD, Reichert TE, Watkins S, Whiteside TL: Fas ligand-positive membranous vesicles isolated from sera of patients with oral cancer induce apoptosis of activated T lymphocytes. Clin Cancer Res. 2005, 11: 1010-1020.

    CAS  PubMed  Google Scholar 

  162. Fourcade O, Simon MF, Viode C, Rugani N, Leballe F, Ragab A, Fournie B, Sarda L, Chap H: Secretory phospholipase A2 generates the novel lipid mediator lysophosphatidic acid in membrane microvesicles shed from activated cells. Cell. 1995, 80: 919-927. 10.1016/0092-8674(95)90295-3.

    Article  CAS  PubMed  Google Scholar 

  163. Willekens FL, Werre JM, Kruijt JK, Roerdinkholder-Stoelwinder B, Groenen-Doepp YA, van den Bos AG, Bosman GJ, Van Berkel TJ: Liver Kupfer cells rapidly remove red blood-derived vesicles from the circulation by scavenger receptors. Blood. 2005, 105: 2141-2145. 10.1182/blood-2004-04-1578.

    Article  CAS  PubMed  Google Scholar 

  164. Pattanapanyasat K, Gonwong S, Chaichompoo P, Noulsri E, Lerdwana S, Sukarpirom K, Siritanaratkul N, Fucharoen S: Activated platelet-derived microparticles in thalassemia. Br J Haematol. 2007, 136: 462-471. 10.1111/j.1365-2141.2006.06449.x.

    Article  CAS  PubMed  Google Scholar 

  165. Sharma R, Muttil P, Yadav AB, Rath SK, Baipai VK, Mani U, Misra A: Uptake of inhalable microparticles affects defence responses of macrophages infected with Mycobacterium tuberculosis H37Ra. J Antimicrob Chemother. 2007, 59: 499-506. 10.1093/jac/dkl533.

    Article  CAS  PubMed  Google Scholar 

  166. Bocci V, Pessina GP, Paulesu L: Studies of factors regulating the aging of human erythrocytes. III. Metabolism and fate of erythrocytic vesicles. Int J Biochem. 1980, 11: 139-142. 10.1016/0020-711X(80)90246-3.

    Article  CAS  PubMed  Google Scholar 

  167. Rank A, Nieuwland R, Crispin A, Gruetzner S, Iberer M, Toth B, Pihusch E: Clearance of platelet microparticles in vivo. Platelets. 2011, 22: 111-116. 10.3109/09537104.2010.520373.

    Article  CAS  PubMed  Google Scholar 

  168. Litvack ML, Post M, Palaniyar N: IgM promotes the clearance of small particles and apoptotic microparticles by macrophages. PLoS One. 2011, 6: e17223-10.1371/journal.pone.0017223.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  169. Al Faraj A, Gazeau F, Wilhelm C, Devue C, Guerin CL, Pechoux C, Paradis V, Clement O, Boulanger CM, Rautou PE: Endothelial cell-derived microparticles loaded with iron oxide nanoparticles: feasibility of MR imaging monitoring in mice. Radiology. 2012, 263: 169-178. 10.1148/radiol.11111329.

    Article  PubMed  Google Scholar 

  170. Caby MP, Lankar D, Vincendeau-Scherrer C, Raposo G, Bonnerot C: Exosomal-like vesicles are present in human blood plasma. Int Immunol. 2005, 17: 879-887. 10.1093/intimm/dxh267.

    Article  CAS  PubMed  Google Scholar 

  171. Toth B, Nikolajek K, Rank A, Nieuwland R, Lohse P, Pihusch V, Friese K, Thaler CJ: Gender-specific and menstrual cycle dependent differences in circulating microparticles. Platelets. 2007, 18: 515-521. 10.1080/09537100701525843.

    Article  CAS  PubMed  Google Scholar 

  172. Grant R, Ansa-Addo E, Stratton D, Antwi-Baffour S, Jorfi S, Kholia S, Krige L, Lange S, Inal J: A filtration-based protocol to isolate human plasma membrane-derived vesicles and exosomes from blood plasma. J Immunol Methods. 2011, 371: 143-151. 10.1016/j.jim.2011.06.024.

    Article  CAS  PubMed  Google Scholar 

  173. Amabile N, Heiss C, Chang V, Angeli FS, Damon L, Rame EJ, McGlothlin D, Grossman W, De Marco T, Yeghiazarians Y: Increased CD62e + endothelial microparticles levels predict poor outcome in pulmonary hypertension patients. J Heart Lung Transplant. 2009, 28: 1081-1086. 10.1016/j.healun.2009.06.005.

    Article  PubMed  Google Scholar 

  174. Nozaki T, Sugiyama S, Sugamura K, Ohba K, Matsuzawa Y, Konishi M, Matsubara J, Akiyama E, Sumida H, Matsui K, Jinnouchi H, Ogawa H: Prognostic value of endothelial microparticles in patients with heart failure. Eur J Heart Fail. 2010, 12: 1223-1228. 10.1093/eurjhf/hfq145.

    Article  PubMed  Google Scholar 

  175. Sinning JM, Losch J, Walenta K, Boehm M, Nickenig G, Werner N: Circulating CD31+/Annexin V + microparticles correlate with cardiovascular outcomes. Eur Heart J. 2011, 32: 2034-2041. 10.1093/eurheartj/ehq478.

    Article  CAS  PubMed  Google Scholar 

  176. Sellam J, Proulle V, Jungel A, Ittah M, Miceli RC, Gottenberg JE, Toti F, Benessiano J, Gay S, Freyssinet JM, Mariette X: Increased levels of circulating microparticles in primary Sjogren’s syndrome, systemic lupus erythematosus and rheumatoid arthritis and relation with disease activity. Arthritis Res Ther. 2009, 11: R156-10.1186/ar2833.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  177. Sheremata WA, Jy W, Delgado S, Minagar A, McLarthy J, Ahn Y: Interferon-beta1a reduces plasma CD31+ endothelial microparticles (CD31 + EMP) in multiple sclerosis. J Neuroinflammation. 2006, 3: 23-10.1186/1742-2094-3-23.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  178. Lowery-Nordberg M, Eaton E, Gonzalez-Toledo E, Harris MK, McGee-Brown J, Ganta CV, Minagar A, Cousineau D, Alexander JS: The effects of high dose interferon-beta1 on plasma microparticles: correlation with MRI parameters. J Neuroinflammation. 2011, 8: 43-10.1186/1742-2094-8-43.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  179. Takikawa M, Nakamura S, Nakamura S, Nambu M, Ishihara M, Fujita M, Kishimoto S, Doumoto T, Yanagibayashi S, Azuma R, Yamamoto N, Kiyosawa T: Enhancement of vascularization and granulation tissue formation by growth factors in human platelet-rich plasma-containing fragmin/protamine microparticles. J Biomed Mater Res B Appl Biomater. 2011, 97: 373-380.

    Article  PubMed  CAS  Google Scholar 

  180. Martinez MC, Andriantsitohaina R: Microparticles in angiogenesis: therapeutic potential. Circ Res. 2011, 109: 110-119. 10.1161/CIRCRESAHA.110.233049.

    Article  CAS  PubMed  Google Scholar 

  181. Morel O, Jesel L, Hugel B, Douchet MP, Zupan M, Chauvin M, Freyssinet JM, 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. 10.1046/j.1538-7836.2003.00010.x.

    Article  CAS  PubMed  Google Scholar 

  182. Nomura S, Omoto S, Yokoi T, Fujita S, Ozasa R, Eguchi N, Shouzu A: Effects of miglitol in platelet-derived microparticle, adiponectin, and selectin level in patients with type 2 diabetes mellitus. Int J Gen Med. 2011, 4: 539-545.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  183. La Vignera S: New immunophenotype of circulating endothelial progenitor cells and endothelial microparticles in patients with erectile dysfunction and metabolic syndrome: effects of taladafil administration. Int Angiol. 2011, 30: 415-423.

    CAS  PubMed  Google Scholar 

  184. Nantakomol D, Dondorp AM, Krudsood S, Udomsangpetch R, Pattanapanyasat K, Combes V, Grau GE, White NJ, Viriyavejakul P, Day NP, Chotivanich K: Circulating red cell-derived microparticles in human malaria. J Infect Dis. 2011, 203: 700-706. 10.1093/infdis/jiq104.

    Article  PubMed Central  PubMed  Google Scholar 

  185. D’Souza-Schorey C, Clancy JW: Tumor-derived microvesicles: shedding light on novel microenvironment modulators and prospective cancer biomarkers. Genes Dev. 2012, 26: 1287-1299. 10.1101/gad.192351.112.

    Article  PubMed Central  PubMed  CAS  Google Scholar 

  186. Ginestra A, Miceli D, Dolo V, Romano FM, Vittorelli ML: Membrane vesicles in ovarian cancer fluids: a new potential marker. Anticancer Res. 1999, 19: 3439-3445.

    CAS  PubMed  Google Scholar 

  187. Baran J, Baj-Krzyworzeka M, Weglarczuk K, Szatanek R, Zembala M, Barbasz J, Czupryna A, Szczepanik A, Zembala M: Circulating tumour-derived microvesicles in plasma of gastric cancer patients. Cancer Immunol Immunother. 2010, 59: 841-850. 10.1007/s00262-009-0808-2.

    Article  CAS  PubMed  Google Scholar 

  188. Coumans FA, Doggen CJ, Attard G, De Bono JS, Terstappen LW: All circulating EpCAM+CK+CD45- objects predict overall survival in castration-resistant prostate cancer. Ann Oncol. 2010, 21: 1851-1857. 10.1093/annonc/mdq030.

    Article  CAS  PubMed  Google Scholar 

  189. Kim HK, Song KS, Park YS, Kang YH, Lee YJ, Lee KR, Kim HK, Ryu KW, Bae JM, Kim S: Elevated levels of circulating platelet microparticles, VEGF, IL-6 and RANTES in patients with gastric cancer: possible role of a metastasis predictor. Eur J Cancer. 2003, 39: 184-191. 10.1016/S0959-8049(02)00596-8.

    Article  CAS  PubMed  Google Scholar 

  190. Langer F, Bokemeyer C: Crosstalk between cancer and haemostasis: implications for cancer biology and cancer-associated thrombosis with focus on tissue factor. Haemostaseologie. 2012, 32: 95-104.

    Article  CAS  Google Scholar 

  191. Nieuwland R, van der Post JA, Lok CA, Kenter G, Sturk A: Microparticles and exosomes in gynecologic neoplasia. Semin Thromb Hemost. 2010, 36: 925-929. 10.1055/s-0030-1267046.

    Article  CAS  PubMed  Google Scholar 

  192. Muradiharan-Chari V, Clancy JW, Sedgwick A, D’Souza-Schorey C: Microvesicles: mediators of extracellular communication during cancer progression. J Cell Sci. 2010, 123: 1603-1611. 10.1242/jcs.064386.

    Article  CAS  Google Scholar 

  193. Buller HR, Van Doormaal FF, Van Sluis GL, Kamphuisen PW: Cancer and thrombosis: from molecular mechanisms to clinical presentations. J Thromb Haemost. 2007, 5: 246-254.

    Article  CAS  PubMed  Google Scholar 

  194. Antonyak MA, Li B, Boroughs LK, Johnson JL, Druso JE, Bryant KL, Holowka DA, Cerione RA: Cancer cell-derived microvesicles induce transformation by transferring tissue transglutaminase and fibronectin to recipient cells. Proc Natl Acad Sci USA. 2011, 108: 4852-4857. 10.1073/pnas.1017667108.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  195. Grange C, Tapparo M, Collino F, Vitillo L, Damasco C, Deregibus MC, Tetta C, Bussolati B, Camussi G: Microvesicles released from human renal cancer stem cells stimulate angiogenesis and formation of lung premetastatic niche. Cancer Res. 2011, 71: 5346-5356. 10.1158/0008-5472.CAN-11-0241.

    Article  CAS  PubMed  Google Scholar 

  196. Shedden K, Xie XT, Chandaroy P, Chang YT, Rosania GR: Expulsion of small molecules in vesicles shed by cancer cells: association with gene expression and chemosensitivity profiles. Cancer Res. 2003, 63: 4331-4337.

    CAS  PubMed  Google Scholar 

  197. Jaiswal R, Gong J, Sambasivam S, Combes V, Mathys JM, Davey R, Grau GE, Bebawy M: Microparticle-associated nucleic acids mediate trait dominance in cancer. FASEB J. 2012, 26: 420-429. 10.1096/fj.11-186817.

    Article  CAS  PubMed  Google Scholar 

  198. Pasguier J, Galas L, Boulange-Lecomte C, Rioult D, Bultelle F, Magal P, Webb G, Le Foll F: Different modalities of intercellular membrane exchanges mediate cell-to-cell P-glycoprotein transfers in MCF-7 breast cancer cells. J Biol Chem. 2012, 287: 7374-7387. 10.1074/jbc.M111.312157.

    Article  CAS  Google Scholar 

  199. Castellana D, Kunzelmann C, Freyssinet JM: Pathophysiologic significance of procoagulant microvesicles in cancer disease and progression. Hamostaseologie. 2009, 29: 51-57.

    CAS  PubMed  Google Scholar 

  200. Janowska-Wieczorek A, Wycoczynski M, Kijowski J, Marguez-Curtis L, Machalinski B, Ratajczak J, Ratajczak MZ: Microvesicles derived from activated platelets induce metastasis and angiogenesis in lung cancer. Int J Cancer. 2005, 113: 752-760. 10.1002/ijc.20657.

    Article  CAS  PubMed  Google Scholar 

  201. Helley D, Banu E, Bouziane A, Banu A, Scotte F, Fischer AM, Oudard S: Platelet microparticles: a potential predictive factor of survival in hormone-refractory prostate cancer patients treated with docetaxel-based chemotherapy. Eur Urol. 2009, 56: 479-484. 10.1016/j.eururo.2008.06.038.

    Article  CAS  PubMed  Google Scholar 

  202. George FD: Microparticles in vascular diseases. Thromb Res. 2008, 122: S55-S59.

    Article  CAS  PubMed  Google Scholar 

  203. Ardoin SP, Shanahan JC, Pisetsky DS: The role of microparticles in inflammation and thrombosis. Scand J Immunol. 2007, 66: 159-165. 10.1111/j.1365-3083.2007.01984.x.

    Article  CAS  PubMed  Google Scholar 

  204. 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. 10.1016/S0002-9440(10)64887-0.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  205. Gruber R, Varga F, Fischer MB, Watzek G: Platelets stimulate proliferation of bone cells: involvement of platelet-derived growth factor, microparticles and membranes. Clin Oral Implant Res. 2002, 13: 529-535. 10.1034/j.1600-0501.2002.130513.x.

    Article  Google Scholar 

  206. 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. 10.1016/j.cardiores.2005.04.007.

    Article  CAS  PubMed  Google Scholar 

  207. Canault M, Leroyer AS, Peiretti F, Leseche G, Tedgui A, Bonardo B, Alessi MC, Boulanger CM, Nalbone G: Microparticles of human atherosclerotic plaques enhance the shedding of tumor necrosis factor-alpha converting enzyme/ADAM 17 substrates, tumor necrosis factor and tumor necrosis factor receptor-1. Am J Pathol. 2007, 171: 1713-1723. 10.2353/ajpath.2007.070021.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  208. Hugel B, Socie G, Vu T, Toti F, Gluckman E, Freyssinet JM, Scrobohaci ML: Elevated levels of circulating procoagulant microparticles in patients with paroxysmal nocturnal hemoglobinuria and aplastic anemia. Blood. 1999, 93: 3451-3456.

    CAS  PubMed  Google Scholar 

  209. Liebman HA, Feinstein DI: Thrombosis in patients with paroxysmal noctural hemoglobinuria is associated with markedly elevated plasma levels of leukocyte-derived tissue factor. Thromb Res. 2003, 111: 235-238. 10.1016/j.thromres.2003.09.018.

    Article  CAS  PubMed  Google Scholar 

  210. Simak J, Holada K, Risitano AM, Zivny JH, Young NS, Vostal JG: Elevated circulating endothelial membrane microparticles in paroxysmal nocturnal haemoglobinuria. Br J Haematol. 2004, 125: 804-813. 10.1111/j.1365-2141.2004.04974.x.

    Article  PubMed  Google Scholar 

  211. Agouni A, Lagrue-Lak-Hal AH, Ducluzeau PH, Mostefai HA, Draunet-Bisson C, Leftheriotis G, Heymes C, Martinez MC, Andriantsithaina R: Endothelial dysfunction caused by circulating microparticles from patients with metabolic syndrome. Am J Pathol. 2008, 173: 1210-1219. 10.2353/ajpath.2008.080228.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  212. Satta N, Toti F, Feugas O, Bohbot A, Dachary-Prigent J, Eschwege V, Hedman H, Freyssinet JM: Monocyte vesiculation is a possible mechanism for dissemination of membrane-associated procoagulant activities and adhesion molecules after stimulation by lipoplysaccharide. J Immunol. 1994, 153: 3245-3255.

    CAS  PubMed  Google Scholar 

  213. Oehmcke S, Moergelin M, Malmstroem J, Linder A, Chew M, Thorlacius H, Herwald H: Stimulation of blood mononuclear cells with bacterial virulence factors leads to the release of pro-coagulant and pro-inflammatory microparticles. Cell Microbiol. 2012, 14: 107-119. 10.1111/j.1462-5822.2011.01705.x.

    Article  CAS  PubMed  Google Scholar 

  214. Perez-Casal M, Thompson V, Downey C, Welters I, Wyncoll D, Thachil J, Toh CH: The clinical and functional relevance of microparticles induced by activated protein C treatment in sepsis. Crit Care. 2011, 15: R195-10.1186/cc10356.

    Article  PubMed Central  PubMed  Google Scholar 

  215. Ogura H, Tanaka H, Koh T, Fujita K, Fujimi S, Nakamori Y, Hosotsubo H, Kuwagata Y, Shimazu T, Sugimoto H: Enhanced production of endothelial microparticles with increased binding to leukocytes in patients with severe systemic inflammatory response syndrome. J Trauma. 2004, 56: 823-830. 10.1097/01.TA.0000084517.39244.46.

    Article  PubMed  Google Scholar 

  216. Stahl AL, Sartz L, Karpman D: Complement activation on platelet-leukocyte complexes and microparticles in enterohemorrahagic Escherichia coli-induced hemolytic uremic syndrome. Blood. 2011, 117: 5503-5513. 10.1182/blood-2010-09-309161.

    Article  PubMed  CAS  Google Scholar 

  217. Mastronardi ML, Mostefai HA, Meziani F, Martinez MC, Asfar P, Andriantsitohaina R: Circulating microparticles from septic shock patients exert differential tissue expression of enzymes related to inflammation and oxidative stress. Crit Care Med. 2011, 39: 1739-1748. 10.1097/CCM.0b013e3182190b4b.

    Article  CAS  PubMed  Google Scholar 

  218. Mortaza S, Martinez CM, Baron-Menguy C, Burban M, de la Bourdonnaye M, Fizanne L, Pierrot M, Cales P, Henrion D, Andriantsitohaina R, Mercat A, Asfar P, Meziani F: Detrimental hemodynamic and inflammatory effects of microparticles originating from septic rats. Crit Care Med. 2009, 37: 2045-2050. 10.1097/CCM.0b013e3181a00629.

    Article  CAS  PubMed  Google Scholar 

  219. Combes V, Taylor TE, Juhan-Vague I, Mege JL, Mwenechanya J, Tembo M, Grau GE, Molyneux ME: Circulating endothelial microparticles in malawian children with severe falciparum malaria complicated with coma. JAMA. 2004, 291: 2542-2544.

    CAS  PubMed  Google Scholar 

  220. Combes V, El-Assaad F, Faille D, Jambou R, Hunt NH, Grau GE: Microvesiculation and cell interactions at the brain-endothelial interface in cerebral malaria pathogenesis. Progr Neurobiol. 2010, 91: 140-151. 10.1016/j.pneurobio.2010.01.007.

    Article  CAS  Google Scholar 

  221. Bhattacharjee S, Van Ooij C, Balu B, Adams JH, Haldar K: Maurer’s clefts of Plasmodium falciparum are secretory organelles that concentrate virulence protein reporters for delivery to the host erythrocyte. Blood. 2008, 111: 2418-2426. 10.1182/blood-2007-09-115279.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  222. Faille D, Combes V, Mitchell AJ, Fontaine A, Juhan-Vague I, Alessi MC, Chimini G, Fusai T, Grau GE: Platelet microparticles: a new player in malaria parasite cytoadherence to human brain endothelium. FASEB J. 2009, 23: 3449-3458. 10.1096/fj.09-135822.

    Article  CAS  PubMed  Google Scholar 

  223. Spycher C, Rug M, Klonis N, Ferguson DJP, Cowman A, Beck H-P, Tilley L: Genesis of and trafficking to the Maurer’s clefts of Plasmodium falciparum infected erythrocytes. Mol Cell Biol. 2006, 26: 4074-4085. 10.1128/MCB.00095-06.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  224. Silverman JM, Clos J, De’Oliveira CC, Shirvani O, Fang Y, Wang C, Foster LJ, Reiner NE: An exosome-based secretion pathway is responsible for protein export from Leishmania and communication with macrophages. J Cell Sci. 2010, 123: 842-852. 10.1242/jcs.056465.

    Article  CAS  PubMed  Google Scholar 

  225. Kuramitsu HK, Kang IC, Qi M: Interactions of Porphyromonas gingivalis with host cells: implications for cardiovascular diseases. J Periodontol. 2003, 74: 85-89. 10.1902/jop.2003.74.1.85.

    Article  PubMed  Google Scholar 

  226. Pussinen PJ, Mattila K: Periodontal infection and atherosclerosis: mere associations?. Curr Opin Lipidol. 2004, 15: 583-588. 10.1097/00041433-200410000-00013.

    Article  CAS  PubMed  Google Scholar 

  227. Nandan D, Tran T, Trinh E, Silverman JM, Lopez M: Identification of leishmania fructose-1,6-biphosphate aldolase as a novel activator of host macrophage Src homology 2 domain containing protein tyrosine phosphatase SHP-1. Biochem Biophys Res Commun. 2007, 364: 601-607. 10.1016/j.bbrc.2007.10.065.

    Article  CAS  PubMed  Google Scholar 

  228. Gomez MA, Contreras I, Halle M, Tremblay ML, McMaster RW, Olivier M: Leishmania GP63 alters host signalling through cleavage-activated protein tyrosine phosphatases. Sci Signal. 2009, 2: ra58-10.1126/scisignal.2000213.

    Article  PubMed  Google Scholar 

  229. McCall LI, Matlashewski G:Localization and induction of the A2 virulence factor inLeishmania: evidence that A2 is a stress response protein. Mol Microbiol. 2010, 77: 518-530. 10.1111/j.1365-2958.2010.07229.x.

    Article  CAS  PubMed  Google Scholar 

  230. Oliveira DL, Freire-de-Lima CG, Nosanchuk JD, Casadevall A, Rodrigues ML, Nimrichter L: Extracellular vesicles from Cryptococcus neoformans modulate macrophage functions. Infect Immun. 2010, 78: 1601-1609. 10.1128/IAI.01171-09.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are thankful to Luke Jasenosky and Aleksandra Gorelova (Harvard University) for editorial help with the manuscript. NSB was supported by a Harvard Pilot Grant and the Immune Disease Institute, and IAV was supported by Russian Foundation for Basic Research grants 11-01517a and 11-01749a.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Natasha S Barteneva.

Additional information

Competing interests

LD is employed by Becton Dickinson Biosciences Inc. Other authors do not have any competing interests.

Authors’ contributions

NSB and IAV wrote the first draft. EFK, MB, JNHS, EDP and LD critically reviewed a manuscript and contributed towards figures. All authors read and approved the final manuscript.

Electronic supplementary material

Additional file 1: Range of MP sizes in different publications.(DOC 42 KB)

Additional file 2: References for Table 1 (Summary of some methods applied for MPs research).(DOC 38 KB)

Additional file 3: MP-based risk stratification of some pathological states.(DOC 54 KB)

12860_2012_660_MOESM4_ESM.doc

Additional file 4: Supplemental Table. Changes in MP levels in peripheral blood of patients in response to treatments.(DOC 45 KB)

Additional file 5: References for Table 2 (MP levels in the plasma of healthy controls).(DOC 34 KB)

Additional file 6: References for Table 3 (MP levels in the plasma and body fluids of patients with cancer).(DOC 42 KB)

12860_2012_660_MOESM7_ESM.doc

Additional file 7: References for Table 4 (MP levels in the plasma and body fluids of patients with different disorders).(DOC 112 KB)

Authors’ original submitted files for images

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Barteneva, N.S., Fasler-Kan, E., Bernimoulin, M. et al. Circulating microparticles: square the circle. BMC Cell Biol 14, 23 (2013). https://doi.org/10.1186/1471-2121-14-23

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2121-14-23

Keywords