Next Article in Journal
Cryptic Diversity and Demographic Expansion of Plasmodium knowlesi Malaria Vectors in Malaysia
Next Article in Special Issue
Liquid Biopsy: A New Avenue for the Diagnosis of Kidney Disease: Diabetic Kidney Disease, Renal Cancer, and IgA Nephropathy
Previous Article in Journal
Inferring Cell–Cell Communications from Spatially Resolved Transcriptomics Data Using a Bayesian Tweedie Model
Previous Article in Special Issue
The Role of PLA2R in Primary Membranous Nephropathy: Do We Still Need a Kidney Biopsy?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Liquid Biopsy at the Frontier of Kidney Diseases: Application of Exosomes in Diagnostics and Therapeutics

1
Department of Internal Medicine, Mayo Clinic, Rochester, MN 55905, USA
2
Department of Biochemistry and Molecular Biology, Mayo Clinic, Rochester, MN 55905, USA
3
Department of Pediatrics, University of Tennessee Health Science Center and Le Bonheur Children’s Hospital, Memphis, TN 38105, USA
4
Children’s Foundation Research Institute, Le Bonheur Children’s Hospital, Memphis, TN 38105, USA
5
Pediatric Medicine Department, St. Jude Children’s Research Hospital, Memphis, TN 38105, USA
*
Author to whom correspondence should be addressed.
Genes 2023, 14(7), 1367; https://doi.org/10.3390/genes14071367
Submission received: 8 May 2023 / Revised: 8 June 2023 / Accepted: 26 June 2023 / Published: 28 June 2023
(This article belongs to the Special Issue Genetic Markers and Liquid Biopsy for Kidney Diseases)

Abstract

:
In the era of precision medicine, liquid biopsy techniques, especially the use of urine analysis, represent a paradigm shift in the identification of biomarkers, with considerable implications for clinical practice in the field of nephrology. In kidney diseases, the use of this non-invasive tool to identify specific and sensitive biomarkers other than plasma creatinine and the glomerular filtration rate is becoming crucial for the diagnosis and assessment of a patient’s condition. In recent years, studies have drawn attention to the importance of exosomes for diagnostic and therapeutic purposes in kidney diseases. Exosomes are nano-sized extracellular vesicles with a lipid bilayer structure, composed of a variety of biologically active substances. In the context of kidney diseases, studies have demonstrated that exosomes are valuable carriers of information and are delivery vectors, rendering them appealing candidates as biomarkers and drug delivery vehicles with beneficial therapeutic outcomes for kidney diseases. This review summarizes the applications of exosomes in kidney diseases, emphasizing the current biomarkers of renal diseases identified from urinary exosomes and the therapeutic applications of exosomes with reference to drug delivery and immunomodulation. Finally, we discuss the challenges encountered when using exosomes for therapeutic purposes and how these may affect its clinical applications.

1. Introduction

In the last decade, precision medicine has emerged as a powerful clinical strategy for various diseases such as kidney diseases, providing a significant improvement in the evolution and outcome of patient management. To date, plasma creatinine, the glomerular filtration rate, and tissue biopsy have represented the standards for the diagnostics of kidney diseases [1,2]. However, these methods present with several complications, such as being insensitive, providing a single snapshot in time of the disease, and being invasive [3]. To overcome the limitations of conventional biopsy, liquid biopsy has emerged as a non-invasive approach for diagnosing and monitoring patients [4]. Liquid biopsy is the process by which biofluids such as blood and urine are used to assess and diagnose diseases [5]. This examination method permits the measurement of biomarkers such as cell-free DNA (cfDNA) and exosomes [6]. Owing to its non-invasiveness, liquid biopsies have the potential to be developed as a screening tool as they allow for the detection and real-time monitoring of disease progression. Urine is the second most commonly used biofluid for clinical diagnostics. Produced by the kidneys, urine serves to eliminate organic and inorganic waste products from the body and, in so doing, helps to maintain body homeostasis. An advantage of working with urine is the fact that it can be collected in a noninvasive manner. However, urine contents and concentrations may vary as a result of factors such as diet, fluid intake, health status, gender, and time of collection, making the data interpretation of urine analysis for diagnostic purposes complicated, hence the need to identify alternative ways of analyzing urine biopsy.
Aside from the standard wastes found in urine, the serendipitous discovery of extracellular vesicles (EVs)/exosomes opened up new possibilities for the use of urine in clinical diagnostics since the EVs/exosomes contain molecules not previously detected in urine [7]. Initially identified in the 1980s and believed to function in eliminating cellular wastes resulting from homeostasis or damaged cells [8], it is now known that exosomes are more than just waste carriers, with the capacity to act as signaling vehicles, delivering complex cargo of nucleic acids [9,10], proteins [11], and lipids [12,13,14] to target cells, eliciting functional responses and promoting phenotypic changes that ultimately reprogram the recipient cells. In addition, the initial characterization of urinary EVs/exosomes identified proteins that were typical of epithelial and urothelial cells of the nephron [15]. As such, exosomes offer prognostic information in a wide range of diseases, such as inflammatory diseases [16,17], tumors [18], and renal diseases [19,20], and play a major role as diagnostic and therapeutic tools by affecting diverse cellular processes such as signal transduction [21], antigen presentation [22], and immune regulation [23,24]. Exosomes generated considerable interest for clinical application as diagnostic and therapeutic tools for several reasons. Their presence in diverse types of biofluids gives them the potential to be identified as biomarkers. In addition, their small size and membrane composition permit them to cross major biological membranes including the blood–brain barrier [25,26,27,28]. Furthermore, their biocompatibility, which reduces immunogenicity, and their lipid bi-layer structure, which protects cargo from degradation, makes them attractive as therapeutic vectors [29].
In recent years, updated technical means in liquid biopsy have made it available for wide application in the clinical diagnosis of a range of diseases. Particularly, isolates from urinary exosomes led to the discovery of biomarkers in diseases such as prostate cancer [30] and bladder cancer [31,32]. Such promising studies, coupled with the fact that urinary EVs/exosomes were considered to originate from cells of the urogenital tract, sparked an exponential growth in exosome research with the goal of discovering EVs/exosome biomarkers for urogenital-tract-related pathologies, providing new possibilities for diagnostic and therapeutic applications of EVs/exosomes in kidney diseases. The diverse cell origins and molecular compositions of urinary exosomes, however, pose analytical difficulties. Therefore, standardizing the methodologies for urine sample collection and EV/exosome isolation would go a long way in the discovery of biomarkers and provide insights that would guide sound clinical decisions. With this viewpoint, this review summarizes advances made in exosome research as it pertains to kidney-related diseases, including acute kidney injury, chronic kidney disease, polycystic kidney disease, and renal cell carcinoma. We give a brief overview on the characterization and isolation of extracellular vesicles. In addition, we discuss urine sample handling and processing and provide recommendations to improve the experimental reproducibility. We further summarize updates on the use of exosomes as a diagnostic and therapeutic tool for targeted kidney diseases, with emphasis on their roles in the identification of biomarkers, drug delivery, and immunomodulation. Finally, we address the current practical challenges encountered with the use and clinical translation of exosomes in kidney diseases. Addressing these issues may provide new mechanistic insights and enable a more sophisticated translation of the use of urinary exosomes as novel biomarkers and therapeutic intervention strategies in kidney diseases.

2. Biogenesis, Composition, Characterization, and Function of Exosomes

2.1. Heterogeneity of Extracellular Vesicles

In the early 1980s, the concept that EVs are formed and released through the budding of the plasma membrane was replaced by the concept that intercellular vesicles mediated by multivesicular bodies (MVBs) are formed and released into the extracellular medium [33,34]. Several years later, the term “exosomes” was coined to refer to these EVs (30–100 nm in diameter) of endosomal origin [8]. Today, based on transmission and immuno-electron microscopy and biochemical methods [35,36], EVs may be broadly divided into two main categories: exosomes and microvesicles. However, the classification of EVs is continuously evolving and newly discovered organelles including migrasomes, generated from migrating cells [37,38], and apoptotic bodies, generated by apoptotic cells [39], are now classified as EVs (Figure 1). It is worth noting that, in addition to apoptotic bodies, dying/apoptotic cells can also release microvesicles and exosomes [40].
EVs in urine originate from different parts of the urogenital tract, including the kidneys, bladder, vaginal tract, and prostate. Urinary EVs may also originate from residing immune cells and bacteria, and also from circulation [41], though it is not clear how these EVs enter the urine. EVs possess unique characteristics, such as their size, morphology, density, presence of marker proteins, and biogenesis (Table 1). For example, exosomes and microvesicles differ in the mechanism of their biogenesis, with exosomes being generated by the inward budding of the endosomal membrane as intraluminal vesicles (ILVs) whereas microvesicles originate by an outward budding at the plasma membrane [42] (Figure 1), and more recently identified from the primary cilium [43,44]. In general, owing to their cellular origin and composition, EVs have been found to play important roles in a variety of cellular processes, such as immune regulation, development, regeneration, and reproduction [45,46,47,48,49], and, as such, are associated to the pathogenesis of a variety of diseases [50]. However, exosomes have received more attention than the other EVs and will be the focus of this review.

2.2. Exosome Biogenesis

The biogenesis of exosomes is strictly regulated such that every step of EV formation is directed by specific components (Figure 2). Exosome biogenesis starts with the formation of an early endosome that matures into a late endosome [61,62]. During endosomal maturation, ILVs are formed via endosomal sorting complex required for transport (ESCRT)-dependent and independent machinery [36,63]. The late ILV-containing endosomes are referred to as multivesicular endosomes (MVEs) or multivesicular bodies (MVBs). The ILVs that eventually become exosomes harbor specific proteins and nucleic acids that are stochastically acquired and/or sorted cytoplasmic and membrane-bound contents originating from processing steps within the donor cell [64,65,66].
Generally, MVBs will fuse with lysosomes to degrade or recycle their contents. The MVBs that do not undergo lysosomal degradation are secreted from the cell via Rab guanosine triphosphatases (Rab-GTPases) [62,67], enhanced by other factors, such as pH levels [68] and calcium concentration [69] within and/or outside the donor cell. The secretion of MVBs is accompanied by the release of exosomes (Figure 2) into the extracellular environment [70], where they may modulate the recipient cell’s activity in an endocrine or paracrine manner [71]. The orientation and distribution of lipids in the bi-layered membrane that result from the invagination of the plasma membrane during the formation of the endosome and subsequent ILVs are crucial to the exosomes’ ability to efficiently mediate cell–cell communications with recipient cells [72]. The lipid bilayer appears to be asymmetrical in the exosome membrane, with phosphatidylserine species typically found in the inner leaflet and sphingomyelin species in the outer leaflet [73]. However, phosphatidylserine is reported to be externalized in apoptotic and malignant cells, acting as an “eat me” signal for macrophages in the immune system [74]. The ordered distribution of lipids in the exosomal membrane, on the other hand, is suggested to be responsible for interactions during exosome formation, release, and delivery to recipient cells. In this regard, studies have revealed that exosomes modify their lipid and metabolite composition depending on the conditions under which they are produced [75,76].

2.3. Exosome Composition

The cellular origin and size of exosomes determine their composition, which may dynamically change as a result of modifications that occur in their parent cell. Exosomes are limited by a bilayered lipid membrane that encapsulates cargo molecules in an inner aqueous core. The lipid composition of exosomes is derived from different sources, which shows that the exosomal lipid membrane is a highly ordered structure enriched in varying lipid classes such as cholesterol and glycosphingolipids [14,77]. This enrichment in lipids is proposed to increase the exosomes’ stability in extracellular environments. The ordered distribution of lipids in the exosomal membrane is believed to play a vital role in the formation of interactions required for exosome biogenesis, release, and delivery/signal transduction to recipient cells [14]. The lipid distribution into exosomes is a dynamic process that responds to several factors and may be modified under different conditions. In one study, for instance, significant variations in the lipid composition of reticulocyte-derived exosomes were observed in response to the physiological changes in the cell during the maturation to erythrocytes, which suggests that lipid sorting for exosome biogenesis could adapt to the cell requirements [78]. In another study, it was demonstrated that prostate cancer PC-3 cells cultured with the ether lipid precursor hexadecylglycerol secreted exosomes enriched in ether lipids, which suggests that external stimuli have the capability to impact the lipid composition of exosomes [75].
The cargo composition of exosomes is unique and complex, differing from one donor cell type to another, the physiological stage of the donor cell, and the fate and function of the exosome [14,78]. Primarily, the cargo consists of peptides, small proteins, and nucleic acids (DNA and RNAs) [11] (Figure 3), all of which are used by the donor cell to transmit signals to other cell populations, coordinate biological functions, and maintain homeostasis. The most frequently identified proteins in exosomes are tetraspanins (e.g., CD9, CD63, and CD81), MVB biogenesis proteins (e.g., Alix and TSG101), membrane transporters and fusion proteins (e.g., GTPases, annexins, and flotillin), heat shock proteins (e.g., HSC70), and lipid-related proteins and phospholipases [11,79]. Several proteins are recognized as specific exosomal markers, among which the tetraspanins CD63 and CD81 are most commonly used [80]. Currently, the exosome content database, Exocarta (http://www.exocarta.org accessed on 16 March 2023), hosts 41,860 protein, >7540 RNA, and 1116 lipid-cataloged entries from 10 distinct species. From these entries, 9769 proteins, 3408 mRNAs, 2838 miRNAs, and 194 lipids have been identified in exosomes from multiple organisms [65,81].
The composition of the exosome provides information about the state of the cell of origin and offers clues regarding its possible effects on the recipient cells [82]. In the kidney, for example, the presence of integrins has been identified on various cell types, including fibroblasts [83] and tubular epithelial cells (TECs) [84,85]. Under physiological conditions, TECs have been found to express αv and β1 integrins. However, under pathological conditions such as kidney injury, they also express β6 in addition to αv and β1 integrins [86]. Similarly, fibroblast cells are known to express α1, α4, α5, and β1 integrins under physiological conditions. However, under pathological conditions associated with fibrosis, they express αv, α5, and β1 integrins [87]. Integrin αv expressed by fibroblast cells has been reported to bind TGF-β, resulting in the stimulation of fibrosis [88], while an increased expression of β1 integrin is associated with epithelial cell polarization [89]. Therefore, one could speculate that, under pathological conditions, exosomes containing αv integrin would stimulate fibrosis in target cells while exosomes containing β1 integrin would enhance TEC polarization and detachment from the basal membrane. It is important to note, however, that the effects of exosomes on target cells may differ due to the varying expressions of cell surface receptors found on different target cells, creating a functional heterogeneity [90]. Likewise, exosome heterogeneity may also arise from differences in the tissue and organ of origin of the exosome.

2.4. Urine Handling for Exosome Research

In urinary exosome studies, the pre-analytical handling of the urine sample is a major source of data variability and can limit reproducibility as preservation and storage methods may impact the outcome [91]. Thus far, there are very limited studies addressing urine sample collection, processing, and storage for exosome research. Nevertheless, pre-analytical variables such as collection (use of preservatives), processing (centrifugation), storage (short-term storage at 4 °C or immediate freezing at −80 °C), and urine handling after thawing (slow or rapid thawing) can alter the results and limit reproducibility. Therefore, consistency in handling urine samples within a study and across research teams is of utmost importance. Due to study specifications, individual research groups have been known to collect, process, and store urine samples differently, making it difficult to come up with a universal standard for urine sample handling in exosome research. In a recent study, however, the authors compared the current methods used for urine collection and storage for urinary EV utilization. They concluded that the addition of protease inhibitors and preservatives such as sodium azide, long-term storage at −80 °C, and extensive vortexing after thawing provides the best quality for urinary EVs [92].

2.5. Exosome Isolation and Characterization

The isolation of pure exosomes is critical to understanding their mode of action and for their potential clinical applications. However, due to their small size and heterogeneity in the exosome population, the isolation of exosomes is challenging. In urine, for example, exosomes originate from different parts of the urogenital tract, including the kidneys, bladder, vaginal tract, and prostate, as well as residing immune cells and bacteria, and also from the circulation [43]. Moreover, within an organ such as the kidney, exosomes can be released from different cell types, such as epithelial and stromal cells. Nevertheless, various techniques have been adopted to facilitate the isolation and characterization of exosomes. Isolation techniques include differential ultracentrifugation [93], size exclusion chromatography [94], immune affinity capture [95,96], exosome precipitation [97], microfluidic-based isolation [98], polymer precipitation [99], and commercially available kits (Table 2). Currently, centrifugation-based techniques are widely considered to be the gold standard for exosome isolation. However, the most essential method for size-based isolation is ultrafiltration [100]. Size exclusion chromatography enables size-based separation on a single column and ensures exosome elution before other soluble components such as proteins [101]. Together with ultrafiltration, size exclusion chromatography techniques are reported to achieve exosome preparations of high purity [102]. It is obvious that each of these isolation techniques has its pros and cons.
Once isolated, exosomes are characterized based on their size, morphology, density, and the presence of marker proteins. However, due to the differences in isolation techniques, heterogeneity of the exosome population, and the difficulty in completely separating cargo profiles, the characterization of exosomes can be particularly challenging. In general, the method used for characterizing exosomes depends on their specific attributes. For size characterization, for example, available methods include nanoparticle tracking analysis (NTA) [118,119] and dynamic light scattering (DLS) [120]. For the characterization of exosome morphology, electron microscopy is the standard method. The morphology observed by scanning electron microscopy (SEM), however, contradicts that of transmission electron microscopy (TEM). While SEM images depicts round-shaped exosomes, TEM images show that exosomes are cup-shaped [118,120]. This difference may be as a result of the type of information provided by the techniques. While SEM provides a three-dimensional image of the surface of the sample, TEM provides a two-dimensional projection image of the inner surface of the sample [121]. It is also important to note that TEM requires very thin sections that are generally less than 150 nm and, in some instances, even less than 30 nm for higher-resolution images [121]. Therefore, sample preparation is quite complex and tedious and may affect exosome properties. For proteomics and biological characterization, mass spectrometry and Western blotting have been widely used.

2.6. Advantages of Exosomes in Liquid Biopsy

Due to their unique features, exosomes have become promising analytes in the field of liquid biopsy. Among the different components such as cfDNA that can be analyzed in liquid biopsy, exosomes are particularly promising because of several reasons. First and foremost, exosomes exist in almost all biofluids and are known to be highly stable owing to their lipid bilayer. This stability can have an effect on the sample storage and transportation, which impacts the clinical applicability of exosomes [122]. Second, exosomes contain the biological information from their parent cells, opening up a valuable avenue for future genetic studies/screening and human disease diagnosis and monitoring. Third, exosomes are relatively easy to identify since they express biological contents such as proteins and lipids that can be used as markers to differentiate them from other extracellular vesicles [123]. Moreover, the specificity of the surface proteins from their parent cells may help to predict the origin of the exosome and provide information on organ-specific disease [124]. Fourth, compared to other liquid biopsies such as circulating tumor cells (CTCs), exosomes are relatively easier to obtain from other biofluids [125]. In addition, it has been reported that a considerable amount of human plasma cfDNA are located in exosomes [126] and that the mitochondrial DNA copy number is higher in exosomes than in plasma in some disease [127]. Fifth, reports indicate that the frequency of detecting mutations in exosomal DNA is higher than that in cfDNA and that exosomal DNA has greater prognostic value compared to cfDNA [128,129]. Finally, exosomes are secreted by living cells, unlike cfDNA, which is secreted during cell death (apoptosis and necrosis). As such, exosomes contain biological information from their parent cells, rendering them more representative and clinically applicable. Undoubtedly, exosomes play an important role in various physiological and pathological processes, and there is evidence to show that exosomes are potential tools for clinical application, including liquid biopsy. Whether there is a difference between urine and blood exosomes as diagnostic biomarkers in kidney disease remains unclear. It has, however, been reported that exosome concentration in urine is substantially lower compared to serum [130]. In addition, exosomes from any biofluid are a representation of the cells in that given microenvironment. Therefore, according to human anatomy and pathophysiology, urinary system diseases would benefit more from urine than blood, suggesting that exosomes from urine would be better diagnostic biomarkers in kidney disease.

3. Exosome-Mediated Mode of Communication

In general, neighboring cells communicate with one another through direct cell–cell contact, such as gap junctions and cell surface protein/protein interactions, while distant cells communicate with one another through secreted soluble factors, such as hormones and cytokines [131,132]. In recent years, exosomes, with their cell-specific cargo, have been recognized as a novel mechanism of intercellular communication [133,134]. Exosomes can mediate cell–cell communication locally and systemically [135]. Once released into the extracellular microenvironment, exosomes can reach the recipient cells and promote phenotypic changes or trigger responses [132,135]. The success of exosomal communication is highly dependent on the effective mechanism of cargo delivery, which may be achieved via receptor–ligand interactions, direct fusion with the plasma membrane of the recipient cell, or via internalization [136] (Figure 4). As such, the exosomes impact recipient cells by the direct stimulation of surface-bound ligands, the transfer of activated receptors to recipient cells, and by epigenetic reprogramming through the delivery of functional epigenetic proteins and miRNAs [137,138,139,140].
Receptor–ligand interactions are a common route for the mediation of immunomodulatory functions [141]. Transmembrane ligands on the exosome surface bind to receptors on the surface of the recipient cells and, depending on the nature of the ligand and receptor, specific downstream signaling cascades are subsequently activated or inhibited [142]. Exosomes also fuse directly to the recipient cell and release their contents into the cytosol [143]. Similar to cell membrane fusion, the lipid bilayer of the exosome partially fuses with that of the recipient cell with the help of Rab GTPases and SNARE family proteins, and eventually expands to form a consistent structure [144]. Exosomes can also undergo internalization by the recipient cell. The process of internalization may occur via various pathways, including clathrin-coated endocytosis, caveolin-mediated endocytosis, lipid-raft-mediated endocytosis, and phagocytosis [143,145,146,147]. However, the fate of the exosomes post-internalization is following the endosomal pathway and eventually fusing with the lysosome for degradation [136,148], or bypassing degradation and being directed to other cellular locations where they mediate functional changes [136].
The role of exosomes in cell–cell communication has been extensively studied and found to have multiple physiological and pathophysiological functions. It is plausible to suggest that these functions are associated with the route of exosome–recipient cell communication. It is also plausible that the mode of communication may be determined by the donor cell type, the exosome membrane composition, and/or the cargoes.

4. Exosomes as Biomarkers for Kidney Diseases

The most commonly used markers for kidney diseases include the estimated glomerular filtration rate (eGFR), blood urea and proteinuria, albuminuria, and serum creatinine. However, these markers are unable to reflect functional changes in the kidney at early disease stages. Identifying novel biomarkers may overcome this limitation and provide a clearer understanding of kidney pathophysiology. In this light, urine contains a considerable number of exosomes that may reflect changes in different cellular compartments. As such, urinary exosomes have been suggested as a promising liquid biopsy for kidney diseases [149]. The composition of urinary exosomes secreted from different segments of the nephron and their relevance to the kidney physiology and pathology of kidney diseases have been investigated. Proteomic analysis of exosomes secreted under physiological and pathological conditions revealed significant changes in protein expression. Cargos released in pathological conditions, such as cancer and inflammation, contain specific constitutive components, such as transmembrane proteins and nucleic acids, which can act as biomarkers for clinical diagnosis, staging disease severity, or assessing therapeutic response [150,151]. Studies from the past few years have identified exosome proteins as biomarkers for different kidney diseases (Table 3). The roles of EVs in renal physiology have been investigated both in vivo and in vitro. EVs are produced and secreted by many cell types in the kidney and proteomic analysis of urinary EVs has demonstrated that the contents of these EVs arise from all segments of the kidney, including proximal tubules, the distal tubule, and the collecting duct [15]. To date, urinary exosomes have been proposed as a promising source of non-invasive biomarkers for the diagnosis and prognosis of various kidney diseases. Urinary exosomes were suggested since they act as indicators of renal function as exosomes secreted by TECs into the urine may vary depending on the physiological and pathological state of the kidney [152,153]. Furthermore, based on the origin, the composition of urinary exosomes may serve as biomarkers for specific kidney diseases [154]. In this section, we discuss the identification of urinary exosomal biomarkers in several common types of kidney diseases, including acute kidney injury, chronic kidney disease, polycystic kidney disease, and renal cell carcinoma.

4.1. Acute Kidney Injury (AKI)

Acute kidney injury (AKI) is a common clinical condition associated with the risk of developing chronic kidney disease (CKD) and end-stage kidney disease (ESKD). Sepsis is the leading cause of AKI in the intensive care unit (ICU) and accounts for 45–70% of all AKI cases [182]. Traditionally, the diagnosis of acute kidney injury (AKI) relies on the serum creatinine and urine output. However, this diagnostic tool is believed to be less sensitive and specific. Since exosomes are secreted by live cells, the detection of urinary exosomes and their contents are explored as potential specific biomarkers for AKI (Table 3). In renal ischemia–reperfusion (I/R) injury, for example, the decreased expression level of aquaporin-1 (AQP1), believed to be controlled by urinary exosomes, is suggested to be a novel urinary biomarker for renal (I/R) injury [155]. In cisplatin-induced AKI, the organic anion transporter 5 (Oat5) was identified as a biomarker for AKI. It was demonstrated that the urinary excretion of exosomal was notably increased, but, when renal injury was ameliorated by the administration of N-acetylcysteine, an Oat5 increase was undetected [156]. Similarly, a significant increase in urine exosome fetuin-A was detected in cisplatin-induced AKI prior to any evidence of morphological injury, suggesting the potential for fetuin-A to serve as a biomarker in AKI patients [183]. In acute tubular necrosis, increased urinary exosome levels of Na+/H+ exchange type-3 were observed and identified as a potential biomarker for AKI [158]. Additionally, urinary exosomal levels of the activating transcription factor 3 (ATF3) were reportedly elevated in AKI patients, even before serum creatinine levels, suggesting that ATF3 as a biomarker may be used for the early diagnosis of AKI [157].
In a sepsis-induced AKI rat model, the microRNAs miR-181a-5p and miR-23b-3p were found to be differentially expressed in circulating extracellular vesicles earlier than creatinine elevation. The expression of these miRNAs may potentially serve as an important tool for the early identification of sepsis-induced AKI and for discriminating sepsis-induced AKI from other causes of AKI [184]. In AKI patients with cirrhosis, the upregulation of maltase–glucoamylase (MGAM), a renal brush border disaccharidase, in urinary exosomes has also been suggested as a potential biomarker, which may differentiate the type of kidney injury in cirrhosis. However, its clinical relevance needs to be further validated [185]. Several other urinary biomarkers have been identified as indicators for the prediction and diagnosis of AKI, including liver fatty-acid-binding protein (L-FABP) [186], neutrophil gelatinase-associated lipocalin (NGAL) [187,188], kidney injury molecule-1 (KIM-1) [189], N-acetyl-β-D-glucosaminidase (NAG) [190], tissue inhibitor metalloproteinase-2 (TIMP-2), insulin growth factor binding protein-7 (IGFBP-7) [191,192], and urinary thioredoxin [193]. Though these biomarkers have been identified, their availability for clinical care has proven slow due to accuracy, the availability of testing platforms, variability in assay techniques, and the cost. It is important to note, however, that IGFBP-7 [194] and thioredoxin [195] are clinically available.

4.2. Chronic Kidney Disease (CKD)

Chronic kidney disease (CKD) is characterized by renal dysfunction, usually diagnosed by the estimated glomerular filtration rate (eGFR) and albuminuria [196]. In recent years, exosomes have been extensively investigated as diagnostic tools in CKD, particularly in renal fibrosis and diabetic nephropathy (Table 3). In renal fibrosis, for example, several studies have identified urinary exosomal micro RNAs in CKD pathology. Urinary exosomal miR-29c [160], miR-181a [162], and miR-200b [163] were reportedly decreased in CKD patients, correlating with the degree of renal fibrosis. By contrast, increased levels of secreting transglutaminase-2 were identified in a unilateral ureteral obstruction (UUO) mouse model [164]. In addition, urinary exosomal miR-21 was increased in CKD patients and inversely correlated with eGFR [197]. Urinary exosomal ceruloplasmin was also found to be increased in CKD rats and patients. Furthermore, immune-reactive ceruloplasmin localized in tubules and collecting ducts of biopsies of CKD patients, while its increased levels were detected in the animals prior to the onset of proteinuria [198]. In more recent years, studies have reported the identification of circular RNA (circRNA) as biomarkers for CKD diseases. The expression level of hsa_circ_0008925 was increased in urinary exosomes from glomerular diseases [165]. In another study, human circRNA analysis reported an increased expression level of hsa_circ_0036649 in urinary exosomes in glomerular diseases [166]. However, its potential role as a biomarker in CKD remains to be further validated.
In addition to renal fibrosis, several proteins and miRNAs have also been identified as biomarkers in diabetic nephropathy. In a proteomic study using label-free comparative techniques to analyze urinary exosomes, it was demonstrated that the urinary exosomal bikunin precursor and histone-lysine N-methyltransferase were increased, whereas voltage-dependent anion-selective channel protein 1 (VDAC1) was decreased in diabetic nephropathy [167]. Additionally, increased levels of AQP-2 and AQP-5 were detected in urinary exosomes derived from diabetic nephropathy patients [168]. These data suggest that these proteins may serve as non-invasive biomarkers for the diagnosis of diabetic nephropathy. Subsequent microarray analysis identified a number of miRNAs as biomarkers for diabetic nephropathy. Increased levels of microRNAs such as miR-371b-5p, miR-320c, miR-572, miR-1234-5p, miR-6068, miR-6133, miR-4270, miR-4739, and miR-638, among others [169], are elevated in urinary exosomes derived from type 2 diabetic nephropathy patients. miR-30d-5p and miR-30e-5p, however, were decreased in type 2 diabetic nephropathy patients [169]. In another study, urinary exosomal levels of several miRNAs, including miR-15b, miR-30a, miR-34a, miR-133b, miR-342, and miR-636, were increased in type 2 diabetic nephropathy patients [170,171]. Additionally, the urinary exosomal level of let-7c-5p was found to be increased, while miR-15b-5p and miR-29c-5p were decreased in type 2 diabetic nephropathy patients [172]. These findings underscore the importance of urinary exosomes and, more particularly, the role of micro RNAs as the source of non-invasive biomarkers for the diagnosis of CKD.

4.3. Polycystic Kidney Disease (PKD)

Polycystic kidney disease (PKD) is the most common inherited kidney disease and is predominantly caused by a mutation in the genes encoding for polycystin-1 (PC-1), polycystin-2 (PC-2), and fibrocystin/polyductin (FCP), all of which are involved in primary cilia structure and function [199]. Proteomic analysis of PKD urinary exosome-like vesicles detected gene products involved in PKD, suggesting that exosomes may play a significant role in primary cilia biology and may serve as a biomarker for PKD. Subsequent studies demonstrated that ADPKD patients with PKD1 gene mutation had decreased levels of polycystin-1 and polycystin-2 but increased level of transmembrane protein 2 in urinary exosomes [15,173,174]. In addition, cystin, the product of the mouse cpk locus, and ADP-ribosylation factor-like 6, the product of the human Bardet–Biedl syndrome gene (BBS3), were found to be abnormally expressed in urinary exosome-like vesicles of patients with PKD [173]. Further proteomic analysis demonstrated increased levels of complement C3 and C9 in urinary extracellular vesicles derived from ADPKD patients with or without progressive CKD and increased levels of envoplakin, periplakin, and villin-1 only in exosomes from ADPKD patients with progressive CKD [175]. This suggests that envoplakin, periplakin, and villin-1 may be used as biomarkers to differentiate between autosomal dominant polycystic kidney disease (ADPKD) patients with or without CKD.
In another study, the activator of G protein signaling 3 (AGS3), involved in the regulation of polycystin ion channel activity, adenylyl cyclase activity, mitotic spindle orientation, and programmed cell death [200,201], was demonstrated to be significantly increased in urinary exosomes from PCK rats and patients [176]. Additionally, AGS3 was suggested to be a beneficial repair protein in tubular epithelial cells [202], which could facilitate the trafficking of proteins to the plasma membrane. This suggests that AGS3 could play a potential role in the trafficking of proteins, enabling exosomal communication between tubular epithelial cells in PKD. In a more recent study, increased urinary exosomal levels of prominin 1 (CD133), the cellular repressor of E1A-stimulated genes 1 (CREG1), and cadherin 4, indicating morphological and proliferation aberrations, were found to be elevated in the exosomes of patients with ADPKD [177]. Notably, increased urinary exosomal CD133 was identified as a potential biomarker that could be used to distinguish medullary sponge kidney disease patients from ADPKD patients [177].
These studies demonstrate that, in PKD, exosome subpopulations with unique compositions trigger diverse biological effects that affect neighboring cells. As a matter of fact, this was reported in a recent study, where the authors demonstrated that cystic-cell-derived extracellular vesicles and urinary exosomes derived from ADPKD patients promoted cyst growth in Pkd1 mutant kidneys and in 3D cultures [19]. Together, these studies draw attention to the key role of exosomes and demonstrate that urinary exosome contents may: (i) serve as biomarkers for PKD, (ii) have the potential to provide a non-invasive method to monitor the progression of PKD, and (iii) provide insight into the biology of tubular epithelial cell function during PKD disease progression.

4.4. Renal Cell Carcinoma (RCC)

Renal cell carcinoma (RCC) is one of the most common cancers, with an incidence of approximately 3.7% of all new cancer cases. As such, it is important to be able to predict the symptoms of the disease before the appearance of symptoms. Comparative proteomics analysis of urinary exosomes from RCC patients and healthy controls identified several urinary exosomal protein markers, amongst which 10 were validated by Western blot analysis, including matrix metalloproteinase-9 (MMP-9), ceruloplasmin (CP), podocalyxin (PODXL), Dickkopf-related protein 4 (DKK 4), carbonic anhydrase IX (CAIX), aquaporin-1 (AQP-1), neprilysin (CD10), dipeptidase 1 (DPEP1), extracellular matrix metalloproteinase inducer (EMMPRIN), and syntenin-1 [178]. Additionally, altered levels of microRNAs have been identified as biomarkers in RCC. In clear cell RCC (ccRCC) patients, for example, increased levels of miR-150-5p and decreased levels of miR-126-3p have been reported. In this same study, it was reported that a combination of urinary miR-126-3p and miR-449a could be used to distinguish between healthy individuals and ccRCC patients with high sensitivity [179]. In another study, urinary exosomal miR-30c-5p and miR-204-5p were identified as potential diagnostic biomarkers for early-stage ccRCC [180,181].
In summary, exosomal contents such as proteins and miRNAs in urine and plasma make it possible to predict the onset and progression of disease. However, their translation to clinical practice faces some challenges. First of all, the isolation methods need to be optimized and standardized to improve reproducibility. Second, the sensitivity and specificity of exosomal components need to be further analyzed and confirmed. Finally, to ensure the accuracy and reliability of the results, more patients need to be enrolled in the clinical trials.

5. Exosomes and Therapeutic Potential

One of the features that make exosomes useful in diagnostics also makes them useful therapeutically. Exosomal surface markers can be used to target specific cell types. Whether for drug delivery or gene therapy, exosomes engineered for a chosen cell type can be packaged with desired components. Stem cell therapy represents a promising new avenue in the treatment of kidney diseases. Several types of stem cells, including induced pluripotent stem cells (iPSCs) and mesenchymal stem cells (MSCs), have been shown to attenuate kidney dysfunction [203]. In recent years, human urine-derived stem cells (USCs) were introduced as a novel promising candidate in the treatment of kidney diseases. Patient-derived USCs are isolated from urine and studies have found that these USCs could be used as a tool to predict the outcome of the kidney disease [204]. USCs are an attractive cell source for a variety of therapies because they are easily accessible, can be consistently produced by the patient, and are relatively free of ethical concerns. USCs exhibit many characteristics of MSCs, and studies have shown that MSCs play an important role in both acute and chronic kidney diseases. Furthermore, treatments with MSCs have entered clinical trials [205]. The following section focuses on the progress of research on the use of exosomes for drug delivery and immunosuppression in kidney diseases, with particular emphasis on MSCs.

5.1. Exosome Mediated Drug Delivery

Exosomes are acknowledged to function in intercellular communication, triggering physiological responses. Once internalized, the exosome contents can be directly released into the cytosol or fuse with the endosomal membrane, evoking an effect on the recipient cell [143]. This inherent ability of exosomes to transfer biochemical materials between cells highlights their potential as drug delivery vehicles. Several studies have investigated the functionality of exosomes in the delivery of materials, including miRNAs [206], siRNAs [207,208,209], and shRNA [210]. Furthermore, their performance as therapeutic vehicles for the delivery of the anti-inflammatory agent curcumin [211,212] and anticancer agents, including paclitaxel [213] and doxorubicin [214], have been investigated. Compared to other delivery systems such as viruses, liposomes, and nanoparticulate systems, known to have undesirable properties, such as immune activation as foreign particles and toxicity, exosomes have been found to increase the loading efficacy of doxorubicin and decrease the adverse effects on major organs, particularly the heart [214,215]. This suggests that doxorubicin delivery via exosomes may alleviate adverse effects, making them a viable vehicle for the chemotherapeutic drug. Furthermore, this suggests that drug delivery via exosomes may be a viable therapeutic tool for the treatment of kidney diseases.

Drug Loading into Exosomes

Considerable advancements have been made in the development and validation of exosome-based drug delivery systems in recent years. Several approaches have been utilized for the loading of exosomes with therapeutic cargoes. The most used approaches may fall under one of the following categories: (i) the ex vitro loading of naïve exosomes isolated from parent cells; (ii) drug loading into parent cells; and (iii) genetically modifying parent cells with DNA encoding therapeutically active compounds (Table 4). In both categories (ii) and (iii), where the parent cells are altered, the drug and therapeutically active compounds are released in the exosomes.
i.
Ex vitro loading of exosomes isolated from parent cells: this is a passive approach of loading exosomes with small molecules (typically lipophilic), nucleic acids, and proteins by co-incubating these compounds with the exosome. The simple incubation of small molecules such as doxorubicin and paclitaxel may easily penetrate the exosome’s membrane. Nucleic acids and proteins, however, require reformation and reshaping techniques such as sonication, electroporation, and elevating incubation temperatures to achieve high loading efficiencies. The ex vitro loading of exosomes approach seems to be the most practical since exosomes obtained from several isolations may be pooled and then loaded with therapeutic cargo.
ii.
Drug loading into parent cells: in this approach, an exogenous compound is loaded into the parent cell and subsequently released in exosomes into the conditioned medium. For example, it has been demonstrated that mesenchymal stem cells (MSCs) treated with paclitaxel secrete exosomes that, in turn, contain paclitaxel [216]. It should be noted, however, that mammalian cells release low quantities of exosomes, and the purification is usually difficult and with a low yield. To circumvent this, bioinspired exosome-mimetic nanovesicles were developed to target and deliver chemotherapeutic drugs [217]. In this study, the nanovesicles were produced by the breaking down of monocytes or macrophages using a serial extrusion technique. Compared to exosomes, these cell-derived nanovesicles have a 100-fold higher production yield. Furthermore, they have a natural targeting ability since they maintain the topology of the plasma membrane proteins of the originating cell [217]. The engineering of parental cells via liposomes has been demonstrated to selectively deliver hydrophobic compounds to the plasma membrane of cancer cells [218]. In this study, the authors introduced synthetic membrane fusogenic liposomes loaded with chemotherapeutic drugs into parent cells. These liposomes were incorporated into the membranes of vesicles/exosomes and then transferred to neighboring cells [218].
iii.
Genetically modified parent cells: in this approach, the genetic material of the parent cell is modified to express, delete, or re-direct the localization of a gene product, and these changes are reflected in the secreted exosomes. For example, a drug delivery system was developed where macrophages were transfected with plasmid DNA encoding different therapeutic proteins for the treatment of neurodegenerative disorders [219]. In another system, adeno-associated virus (AAV) capsids were incorporated into extracellular microvesicles termed “vexosomes” (microvesicle-associated vectors). In this study, the authors found that, during the production of AAV vectors, a fraction of released vectors were associated with vexosomes, which enhanced gene transfer in cultured cells compared to conventionally purified AAV [220].
Table 4. Methods of exosomal bioengineering and their advantages and disadvantages.
Table 4. Methods of exosomal bioengineering and their advantages and disadvantages.
Loading ApproachFunctionsAdvantagesDisadvantagesReferences
Passive loadingSimple incubation of exosomes with drugs.Incorporation of lipophilic drugs.
-
Easy to perform. Does not affect exosome integrity.
Low loading efficiency.[211,216]
Simple incubation of parent cell with drugs.Incorporation of lipophilic drugs.
-
Easy to perform.
-
Does not affect exosome integrity.
-
Low loading efficiency.
-
May cause cytotoxicity to parent cell.
[216,221]
Active loadingSonication.Incorporation of drugs, proteins, and peptides.High loading efficiency.Compromises exosome membrane integrity.[222]
Electroporation.Incorporation of drugs and large molecules such as nucleic acids and peptides.Loading efficiency is high.May cause RNA aggregation and exosome instability[223,224]
Extrusion.Incorporation of drugs.High loading efficiency.High loading efficiency.[222,225]
Freeze–thaw. Incorporation of proteins and peptides.Medium loading efficiency.
-
May inactivate cargo.
-
Aggregates may be formed.
[222]
Use of membrane permeabilizers (surfactants).Incorporation of proteins and peptides.High loading efficiency.
-
Compromises exosome integrity.
-
May inactivate cargo.
[222,225]
Transfection. Incorporation of nucleic acids, proteins, and peptides.High loading efficiency and molecular stability.Toxicity of transfection agents.[222,226]
Exosome-based drug formulations may be applied to a wide variety of disorders. However, the choice of loading approach may be dictated by the type of cargo to be loaded, the conditions appropriate for the loading of any given type of cargo, and the site of the disease. Though each approach has its advantages and limitations as summarized in the table above (Table 4), a series of them have been successfully used for exosome-mediated drug delivery.

5.2. Exosome-Mediated Drug Delivery in Kidney Diseases

For kidney-related diseases, identifying a prime target for potential exosome-based therapy could have significant implications for therapeutic purposes being that the kidney is made up of several cell types with distinct functions. As such, understanding exosome production, the molecules that they carry, and their interactions within the kidney could have far-reaching implications for therapeutic purposes of kidney diseases. Studies have demonstrated that the purification and use of exosomes from particular cells could be a promising approach for the therapeutic purposes of kidney diseases. An example of this is the isolation and use of exosomes derived from MSCs. MSCs are commonly used in cellular therapy trials for regenerative medicine and immunomodulation [227,228]. Recently, MSCs have been shown to release exosomes, which may function as paracrine mediators between MSCs and target cells [229,230]. Furthermore, MSC-derived exosomes (MSC-Exos) have been found to recapitulate the biological activity of MSCs; as such, they may serve as an alternative therapy to MSCs [231]. Reports indicate that MSC-derived exosomes are ideal vehicles for the transportation and delivery of molecules including therapeutic genes, drugs, and RNA to targeted cells [232].
In the last decade, the use of exosomes in drug delivery has been considerably studied in AKI and has proven to be a promising strategy for the treatment. In one study, MSCs engineered to overexpress miRNA-let7c selectively targeted damaged kidneys and this upregulated miR-let7c-MSC therapy, attenuated kidney injury, and significantly decreased fibrosis by downregulating fibrosis markers, including collagen IVα1, metalloproteinase-9, transforming growth factor (TGF)-β1, and TGF-β type 1 receptor (TGF-βR1), in UUO mouse kidneys compared with non-targeted control MSCs [233]. In another study of I/R-induced injury in AKI, exosomes derived from MSCs pre-conditioned with melatonin were found to provide the best protective effect against kidney injury compared to therapy by MSCs or exosomes derived from non-preconditioned MSCs. This was demonstrated by a decreased kidney injury, reduced blood levels of kidney damage markers, declined oxidative stress status, inhibition of inflammation, and improved regeneration [234]. In another study, exosomes derived from melatonin-treated healthy MSCs were used to determine their therapeutic potential in CKD [235]. Treatment with melatonin increased the expression of cellular prion protein through the upregulation of miR-4516 in MSC-Exos. This study suggested that the treatment of CKD-MSCs with melatonin exosomes might be a strategy for the development of autologous MSC-based therapeutics for patients with CKD [235]. In another study, dexamethasone and glucocorticoid receptors were delivered to the kidneys by macrophage-cell-derived microvesicles expressing integrin surface markers. This attenuated renal injury with an enhanced therapeutic efficacy against renal inflammation and fibrosis in murine models of LPS- or adriamycin-induced nephropathy [236]. Other than biological components such as miRNAs, biochemical materials engineered to function with exosomes have also been reported to promote therapeutic efficacy in AKI. In one study, hydrogels containing Arg-Gly-Asp (RGD) peptide were produced and were functionally determined to sustain the retention and stability of EVs and to increase EV affinity. MSC-derived extracellular vesicles–RGD hydrogels facilitated MSC-derived let-7a-5p-containing EVs, improving therapeutic efficacy in renal repair in AKI [237].
In spite of the essential role that exosomes play in kidney disease pathology, their effectiveness still needs to be investigated. In addition, due to the low yield of exosomes, the research field of engineered exosomes is starting to be concentrated on the function of cell-derived nanovesicles [217,238]. So far, these studies demonstrate that engineered exosomes show beneficial effects on kidney injury. Furthermore, these studies have determined that loading other protective molecules in combination with biological materials or particles enhances protective effects toward kidney diseases and is worth further studies.

5.3. Exosome-Mediated Immunomodulation in Kidney Diseases

In addition to their aforementioned role as biomarkers and in drug delivery, exosomes are capable of influencing multiple aspects of the immune system. A number of studies have reported the vital role of exosomes in immunomodulation and immunosuppression. Amongst these studies, the role of exosomes in the regulation of inflammation in kidney injury has been extensively studied. Previous studies have pointed out that macrophages play important roles in the progress of AKI and AKI to CKD transition [239]. Macrophages adopt two distinct polarization states, switch between pro-inflammatory macrophages (M1s) or anti-inflammatory macrophages (M2s) activation phenotypes in response to varied external stimuli, and make different contributions to disease at different stages [240]. Exosomes derived from MSCs, multipotent cells with beneficial roles in tissue repair, have been found to contribute to immunomodulatory effects by regulating inflammation. For example, bone marrow MSC-derived exosomes (MSC-Exo) were reported to accelerate renal self-repair in ischemia–reperfusion injury (IRI)-induced mice by reducing M1 macrophage polarization and activating M2 macrophages polarization, thereby decreasing the levels of pro-inflammatory cytokines IL-1β, IL-6, and TNF-α while increasing the anti-inflammation factor IL-10. These anti-inflammatory effects were further amplified by indoleamine 2,3-dioxygenase overexpression in MSCs [241]. In addition to their role in immunomodulation, MSC-Exo has been reported to have a renoprotective function by regulating processes including cell proliferation, fibrosis, oxidative stress, autophagy, apoptosis, and necrosis (Table 5).
Table 5. The mechanisms of MSC-derived exosomes in alleviating kidney injury.
Table 5. The mechanisms of MSC-derived exosomes in alleviating kidney injury.
Source of ExosomeBiological MaterialKidney DiseaseConclusionsReferences
mBMMSCsCCR2I/R injury—AKISuppressed CCL2 activity, alleviated inflammation.[242]
mADSCsmiR-486Diabetic nephropathyInhibited Smad1/mTOR signaling pathway, increased autophagy, and reduced podocyte apoptosis.[243]
hWJMSCsmiR-15a/b, miR-16I/R injury—AKIDecreased CX3CL1 expression, alleviated inflammation.[244]
miR-30I/R injury—AKISuppressed DRP1 and mitochondrial fragmentation, showed anti-apoptotic effects.[245]
hP-MSCsmiR-200a-3pI/R injury—AKIActivated the Keap1-Nrf2 signaling pathway and exerted antioxidant effects.[246]
hUSCsmiR-146a-5pI/R injury—AKIDegraded IRAK1 and inhibited NF-κB signaling pathway.[247]
miR-216a-5pI/R injury—AKIDownregulated PTEN, anti-apoptotic effect on HK-2 cells[248]
Abbreviations used: mBMMSCs—mouse bone marrow mesenchymal stem cells; CCR2—C-C motif chemokine receptors 2; CCL2—C-C motif chemokine ligand 2; mADSCs—mouse adipose-derived stem cells; CX3CL1—C-X3-C motif chemokine ligand 1; DRP1—dynamin-related protein 1; hWJMSCs—human Wharton jelly mesenchymal stromal cells; hP-MSCs—human placenta-derived mesenchymal stem cells; hUSCs—human urine-derived stem cells; Keap1—Kelch-like ECH-associated protein 1; Nrf2—nuclear factor erythroid 2-related factor 2; IRAK1—degraded targeted and degraded the 3′UTR of interleukin-1 receptor-associated kinase 1 mRNA; PTEN—phosphatase and tensin homolog; HK-2—human proximal tubular epithelial cells.
Several studies have demonstrated that miRNAs contained within MSC-Exo play a role in immunosuppression. In one study, it was demonstrated that let-7b, a member of the let-7 miR family, contained in MSC-Exo induced the generation of the M2 immunosuppressive phenotype in renal macrophages. In addition, mice that received let-7b containing MSC-Exo showed lower concentrations of M1-derived pro-inflammatory cytokines TNF-α and IL-1β measured in I/R-injured kidneys [249]. In another study, it was demonstrated that miR-21 had anti-inflammatory properties and contributed to alleviating I/R-induced AKI [250]. In this study, the authors reported that miR-21 contained in MSC-Exo reduced NF-κB activity in the renal infiltration and maturation of dendritic cells (DCs). The administration of miR-21 containing MSC-Exo attenuated the production of pro-inflammatory cytokines such as IL-12, IL-6, and TNF-α, and reduced the activation of Th1 and Th17 cell-driven inflammation in I/R-injured kidneys [250].
In addition to MSC-Exo, miRNA-containing exosomes derived from renal TECs have been found to play a significant role in modulating inflammation. In a recent study, exosomal miR-19b-3p released from renal TECs was demonstrated to mediate the cross-talk between renal TECs and macrophages, which subsequently induced tubulointerstitial inflammation [251]. The authors found that miR-19b-3p was notably increased in TEC-derived exosomes from an LPS-induced AKI mouse model and in an adriamycin-induced chronic proteinuric kidney disease mouse model. They further reported that TEC-derived exosomal miR-19b-3p was internalized by macrophages, leading to M1 macrophage activation through the NF-κB/SOCS-1 pathway, thereby promoting tubulointerstitial inflammation. The administration of TEC-derived exosomes with miR-19b-3p inhibition significantly decreased the expression levels of IL-6, MCP-1, and iNOS mRNA in recipient macrophages and resulted in decreased renal inflammation [251].
Apart from the roles of MSC-Exos and TEC-derived exosomes on immunomodulation and immunosuppression in kidney injury, reports have shown that tumor-derived exosomes play a significant role in tumor progression by mediating the cross-talk between tumor cells and immune cells such as macrophages. In a recent study, it was demonstrated that RCC-derived exosomes containing a high amount of a new long non-coding RNA (lncRNA), named lncRNA activated in RCC with sunitinib resistance (lncARSR), played an important role in modulating inflammation [252]. LncARSR was found to interact with miR-34/miR- 449 to activate the STAT3 pathway, resulting in the polarization of the M1 to M2 macrophage phenotype in RCC cells. This polarization resulted in the secretion of anti-inflammatory factors such as IL-10, creating a more suitable microenvironment for tumor metastasis [252]. Sunitinib is a first-line targeted drug for metastatic RCC; however, its resistance is a major challenge for RCC patients. LncARSR correlated with a clinically poor sunitinib response and was found to promote sunitinib resistance by binding to miR-34/miR-449 and modulating the expression of AXL and c-MET in RCC cells [253]. Furthermore, resistant cells could secrete lncARSR via exosomes, transforming sunitinib-sensitive cells into resistant cells, thereby disseminating drug resistance. Targeting lncARSR was found to restore the sunitinib response, suggesting that lncARSR may serve as a predictor and a potential therapeutic target for sunitinib resistance in RCC [253].
Increasing evidence has demonstrated that exosome-containing miRNA derived from MSCs and TECs play a pivotal role in the immunomodulation and immunosuppression of kidney diseases, such as AKI and RCC. In addition, among the mechanisms by which these exosome-derived miRNA function, there is considerable evidence demonstrating that exosome-derived miRNAs function through macrophage polarization from M1 to M2 phenotypes. Taken together, these studies suggest that genetically modifying or engineering MSCs and TECs for the targeted delivery of miRNA may pave the way for an effective approach to improving kidney disease therapy.

6. Summary and Future Perspectives

Liquid biopsy has opened up a new avenue for the early diagnosis and treatment of kidney diseases, especially when tissue samples are difficult to or cannot be obtained. Using liquid biopsy and circulating biomarkers in kidney diseases allows for the assessment of molecular profiles of specific nephron cell. These circulating biomarkers could be used as non-invasive tools to be able to evaluate disease progression and the response to on-going treatments, thereby guiding the medical choice for further treatment strategies. However, there is the need for advanced molecular detection methods to enhance the functionality and reliability of liquid biopsy for clinical practice.
The potential of exosomes to reflect changes in the different segments of the nephron and their release in the urine emphasizes their role in the pathophysiology of kidney diseases. In the field of kidney biology and disease, the therapeutic applications of exosomes thus far comprise their use as prognostic and diagnostic biomarkers, drug delivery carriers, and cell-free therapy. Studies have particularly focused on the potential of exosomes as biomarkers, targeting primarily urinary exosomes, given the ease of collection and ability for urine to serve as a liquid biopsy. Several promising exosome-derived biomarkers have demonstrated the ability to detect early kidney damage and localized kidney injury and to predict disease progression and severity. Increasing evidence has also demonstrated the extraordinary therapeutic potential of MSCs in various kidney diseases, including ischemic diseases such as AKI. Currently, exosomes derived from various sources, including MSCs and TECs, have been reported to modulate several processes, including inflammation, fibrosis, and oxidative stress, which subsequently alleviates renal I/R injury and exerts reparative effects on the kidney. This suggests that MSC-derived and TEC-derived exosomes may serve as promising therapeutic candidates for ischemic AKI. Together, these studies demonstrate the potential of exosomes to improve the diagnostic and clinical treatment of kidney diseases.
Despite the advances outlined in this article, it should be noted that the translation of exosomes into clinical practice for kidney diseases is lacking. In a large number of preclinical studies, exosome-based therapeutics have shown significant efficacy in various kidney diseases. However, their clinical efficacy has not been up to par, with only a couple of clinical trials aiming to explore the clinical efficacy of exosomes against RCC. This suggests that there are still many challenges ahead and questions that need to be addressed. Some of the challenges that need to be resolved are the lack of standardization and consistency in exosome detection, isolation, and purification methods, low exosome yield, which cell type to use for exosome derivation, and the low stability and retention of MSC-derived exosomes. As it pertains to the exosome yield, previous studies have reported that the large-scale manufacturing of exosomes can be obtained via rapid purification. However, the viability of this technique still needs to be assessed with different cell types. Furthermore, the concept of utilizing exosomes as a drug delivery vehicle is a fast-growing area in kidney research. While drug-loaded exosomes may serve as a next-generation drug delivery system, they lack certain features to qualify as vehicles for drug delivery, such as their loading capacity. Exosomes need to be able to hold a considerable quantity of therapeutic cargoes to be able to deliver desirable therapeutic effects. However, since they already contain cargo, their loading capacity is low.
Aside from their inadequacy in their drug loading capacity, the targeting of therapeutic exosomes to specific kidney cells is essential. Fortunately, advanced biotechnological studies in exosome engineering with the goal of designing highly specialized exosomes capable of site-specific targeting to particular tissue or cell types and loading appropriate bioactive cargo into the lumen or the surface of the exosome is ongoing. As research for the development of more viable exosomes for therapeutic purposes in kidney diseases is ongoing, it may be helpful to take into account fundamental questions that remain unanswered. What are the physiological and pathological stimuli that lead to the production, release, and uptake of exosomes by different renal cells? What mechanisms determine the specificity of exosome targets? How does the kidney disease condition affect the composition and the levels of urinary exosomes? Addressing these questions may provide new mechanistic insights and set the stage for the sophisticated development and subsequent translation of exosomes as therapeutic intervention strategies in kidney diseases.
In conclusion, with the rising incidence of kidney diseases, the need for novel prognostic, diagnostic, and therapeutic strategies have never been greater. Studies up to date demonstrate that exosome-based clinical applications hold promise as a next-generation treatment trajectory for kidney diseases. However, the mass production of exosomes, selection of exosome donor cells, drug loading methods, and administration routes are all prominent issues that need to be resolved. With the continuous advancements in biotechnology, we expect these limitations to be overcome. In the end, exosome-based prognosis, diagnosis, and therapy have promising and important clinical significance in the field of nephrology.

Funding

X. Li acknowledges support from the National Institutes of Health grant R01 DK129241 and DK126662. E. Agborbesong acknowledges support from the National Institutes of Health T32 grant DK007013.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Levey, A.S.; Inker, L.A.; Coresh, J. GFR Estimation: From Physiology to Public Health. Am. J. Kidney Dis. 2014, 63, 820–834. [Google Scholar] [CrossRef] [Green Version]
  2. Dhaun, N.; Bellamy, C.O.; Cattran, D.C.; Kluth, D.C. Utility of renal biopsy in the clinical management of renal disease. Kidney Int. 2014, 86, 1268. [Google Scholar] [CrossRef] [Green Version]
  3. De Oliveira, B.D.; Xu, K.; Shen, T.H.; Callahan, M.; Kiryluk, K.; D’Agati, V.D.; Tatonetti, N.P.; Barasch, J.; Devarajan, P. Molecular nephrology: Types of acute tubular injury. Nat. Rev. Nephrol. 2019, 15, 599–612. [Google Scholar] [CrossRef]
  4. Ignatiadis, M.; Sledge, G.W.; Jeffrey, S.S. Liquid biopsy enters the clinic—implementation issues and future challenges. Nat. Rev. Clin. Oncol. 2021, 18, 297–312. [Google Scholar] [CrossRef] [PubMed]
  5. Macias, M.; Alegre, E.; Diaz-Lagares, A.; Patino, A.; Perez-Gracia, J.L.; Sanmamed, M.; Lopez-Lopez, R.; Varo, N.; Gonzalez, A. Liquid Biopsy: From Basic Research to Clinical Practice. Adv. Clin. Chem. 2018, 83, 73–119. [Google Scholar] [CrossRef] [PubMed]
  6. Burnham, P.; Dadhania, D.; Heyang, M.; Chen, F.; Westblade, L.F.; Suthanthiran, M.; Lee, J.R.; De Vlaminck, I. Urinary cell-free DNA is a versatile analyte for monitoring infections of the urinary tract. Nat. Commun. 2018, 9, 2412. [Google Scholar] [CrossRef] [Green Version]
  7. Santucci, L.; Candiano, G.; Petretto, A.; Bruschi, M.; Lavarello, C.; Inglese, E.; Righetti, P.G.; Ghiggeri, G.M. From hundreds to thousands: Widening the normal human Urinome (1). J. Proteom. 2015, 112, 53–62. [Google Scholar] [CrossRef] [Green Version]
  8. Johnstone, R.M.; Adam, M.; Hammond, J.R.; Orr, L.; Turbide, C. Vesicle formation during reticulocyte maturation. Association of plasma membrane activities with released vesicles (exosomes). J. Biol. Chem. 1987, 262, 9412–9420. [Google Scholar] [CrossRef]
  9. Valadi, H.; Ekstrom, K.; Bossios, A.; Sjostrand, M.; Lee, J.J.; Lotvall, J.O. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell. Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef] [Green Version]
  10. Waldenstrom, A.; Genneback, N.; Hellman, U.; Ronquist, G. Cardiomyocyte microvesicles contain DNA/RNA and convey biological messages to target cells. PLoS ONE 2012, 7, e34653. [Google Scholar] [CrossRef] [Green Version]
  11. Simpson, R.J.; Lim, J.W.; Moritz, R.L.; Mathivanan, S. Exosomes: Proteomic insights and diagnostic potential. Expert. Rev. Proteom. 2009, 6, 267–283. [Google Scholar] [CrossRef]
  12. Laulagnier, K.; Motta, C.; Hamdi, S.; Roy, S.; Fauvelle, F.; Pageaux, J.F.; Kobayashi, T.; Salles, J.P.; Perret, B.; Bonnerot, C.; et al. Mast cell- and dendritic cell-derived exosomes display a specific lipid composition and an unusual membrane organization. Biochem. J. 2004, 380, 161–171. [Google Scholar] [CrossRef]
  13. Trajkovic, K.; Hsu, C.; Chiantia, S.; Rajendran, L.; Wenzel, D.; Wieland, F.; Schwille, P.; Brugger, B.; Simons, M. Ceramide triggers budding of exosome vesicles into multivesicular Endosomes. Science 2008, 319, 1244–1247. [Google Scholar] [CrossRef] [PubMed]
  14. Llorente, A.; Skotland, T.; Sylvanne, T.; Kauhanen, D.; Rog, T.; Orlowski, A.; Vattulainen, I.; Ekroos, K.; Sandvig, K. Molecular lipidomics of exosomes released by PC-3 prostate cancer cells. Biochim. Biophys. Acta 2013, 1831, 1302–1309. [Google Scholar] [CrossRef] [PubMed]
  15. Pisitkun, T.; Shen, R.F.; Knepper, M.A. Identification and proteomic profiling of exosomes in human urine. Proc. Natl. Acad. Sci. USA 2004, 101, 13368–13373. [Google Scholar] [CrossRef] [Green Version]
  16. Lasser, C.; O’Neil, S.E.; Shelke, G.V.; Sihlbom, C.; Hansson, S.F.; Gho, Y.S.; Lundback, B.; Lotvall, J. Exosomes in the nose induce immune cell trafficking and harbour an altered protein cargo in chronic airway inflammation. J. Transl. Med. 2016, 14, 181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Zhang, H.; Wang, L.; Li, C.; Yu, Y.; Yi, Y.; Wang, J.; Chen, D. Exosome-Induced Regulation in Inflammatory Bowel Disease. Front. Immunol. 2019, 10, 1464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Gluszko, A.; Mirza, S.M.; Piszczatowska, K.; Kantor, I.; Struga, M.; Szczepanski, M.J. The role of tumor-derived exosomes in tumor angiogenesis and tumor progression. Curr. Issues Pharm. Med. Sci. 2019, 32, 193–202. [Google Scholar] [CrossRef] [Green Version]
  19. Ding, H.; Li, L.X.; Harris, P.C.; Yang, J.; Li, X. Extracellular vesicles and exosomes generated from cystic renal epithelial cells promote cyst growth in autosomal dominant polycystic kidney disease. Nat. Commun. 2021, 12, 4548. [Google Scholar] [CrossRef] [PubMed]
  20. Thongboonkerd, V. Roles for Exosome in Various Kidney Diseases and Disorders. Front. Pharmacol. 2020, 10, 1655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Gangoda, L.; Boukouris, S.; Liem, M.; Kalra, H.; Mathivanan, S. Extracellular vesicles including exosomes are mediators of signal transduction: Are they protective or pathogenic? Proteomics 2015, 15, 260–271. [Google Scholar] [CrossRef] [Green Version]
  22. Mittelbrunn, M.; Gutierrez-Vazquez, C.; Villarroya-Beltri, C.; Gonzalez, S.; Sanchez-Cabo, F.; Gonzalez, M.A.; Bernad, A.; Sanchez-Madrid, F. Unidirectional transfer of microRNA-loaded exosomes from T cells to antigen-presenting cells. Nat. Commun. 2011, 2, 282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Poggio, M.; Hu, T.; Pai, C.C.; Chu, B.; Belair, C.D.; Chang, A.; Montabana, E.; Lang, U.E.; Fu, Q.; Fong, L.; et al. Suppression of Exosomal PD-L1 Induces Systemic Anti-tumor Immunity and Memory. Cell 2019, 177, 414–427.e13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Wang, S.; Li, F.; Ye, T.; Wang, J.; Lyu, C.; Qing, S.; Ding, Z.; Gao, X.; Jia, R.; Yu, D.; et al. Macrophage-tumor chimeric exosomes accumulate in lymph node and tumor to activate the immune response and the tumor microenvironment. Sci. Transl. Med. 2021, 13, eabb6981. [Google Scholar] [CrossRef]
  25. Caby, M.P.; Lankar, D.; Vincendeau-Scherrer, C.; Raposo, G.; Bonnerot, C. Exosomal-like vesicles are present in human blood plasma. Int. Immunol. 2005, 17, 879–887. [Google Scholar] [CrossRef] [Green Version]
  26. Keller, S.; Rupp, C.; Stoeck, A.; Runz, S.; Fogel, M.; Lugert, S.; Hager, H.D.; Abdel-Bakky, M.S.; Gutwein, P.; Altevogt, P. CD24 is a marker of exosomes secreted into urine and amniotic fluid. Kidney Int. 2007, 72, 1095–1102. [Google Scholar] [CrossRef] [Green Version]
  27. Asea, A.; Jean-Pierre, C.; Kaur, P.; Rao, P.; Linhares, I.M.; Skupski, D.; Witkin, S.S. Heat shock protein-containing exosomes in mid-trimester amniotic fluids. J. Reprod. Immunol. 2008, 79, 12–17. [Google Scholar] [CrossRef]
  28. Street, J.M.; Barran, P.E.; Mackay, C.L.; Weidt, S.; Balmforth, C.; Walsh, T.S.; Chalmers, R.T.A.; Webb, D.J.; Dear, J.W. Identification and proteomic profiling of exosomes in human cerebrospinal fluid. J. Transl. Med. 2012, 10, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Wu, P.P.; Zhang, B.; Ocansey, D.K.W.; Xu, W.; Qian, H. Extracellular vesicles: A bright star of nanomedicine. Biomaterials 2021, 269, 120467. [Google Scholar] [CrossRef]
  30. Clos-Garcia, M.; Loizaga-Iriarte, A.; Zuniga-Garcia, P.; Sanchez-Mosquera, P.; Cortazar, A.R.; Gonzalez, E.; Torrano, V.; Alonso, C.; Perez-Cormenzana, M.; Ugalde-Olano, A.; et al. Metabolic alterations in urine extracellular vesicles are associated to prostate cancer pathogenesis and progression. J. Extracell. Vesicles 2018, 7, 1470442. [Google Scholar] [CrossRef] [Green Version]
  31. Zhan, Y.; Du, L.T.; Wang, L.S.; Jiang, X.M.; Zhang, S.J.; Li, J.; Yan, K.Q.; Duan, W.L.; Zhao, Y.H.; Wang, L.L.; et al. Expression signatures of exosomal long non-coding RNAs in urine serve as novel non-invasive biomarkers for diagnosis and recurrence prediction of bladder cancer. Mol. Cancer 2018, 17, 142. [Google Scholar] [CrossRef]
  32. Lee, J.; McKinney, K.Q.; Pavlopoulos, A.J.; Niu, M.; Kang, J.W.; Oh, J.W.; Kim, K.P.; Hwang, S. Altered Proteome of Extracellular Vesicles Derived from Bladder Cancer Patients Urine. Mol. Cells 2018, 41, 179–187. [Google Scholar] [CrossRef]
  33. Harding, C.; Heuser, J.; Stahl, P. Receptor-mediated endocytosis of transferrin and recycling of the transferrin receptor in rat reticulocytes. J. Cell. Biol. 1983, 97, 329–339. [Google Scholar] [CrossRef] [PubMed]
  34. Pan, B.T.; Johnstone, R.M. Fate of the transferrin receptor during maturation of sheep reticulocytes in vitro: Selective externalization of the receptor. Cell 1983, 33, 967–978. [Google Scholar] [CrossRef] [PubMed]
  35. Thery, C.; Amigorena, S.; Raposo, G.; Clayton, A. Isolation and characterization of exosomes from cell culture supernatants and biological fluids. Curr. Protoc. Cell. Biol. 2006, 3, 22. [Google Scholar] [CrossRef]
  36. Raposo, G.; Stoorvogel, W. Extracellular vesicles: Exosomes, microvesicles, and friends. J. Cell. Biol. 2013, 200, 373–383. [Google Scholar] [CrossRef] [Green Version]
  37. Ma, L.; Li, Y.; Peng, J.Y.; Wu, D.; Zhao, X.X.; Cui, Y.T.; Chen, L.L.; Yan, X.J.; Du, Y.N.; Yu, L. Discovery of the migrasome, an organelle mediating release of cytoplasmic contents during cell migration. Cell Res. 2015, 25, 24–38. [Google Scholar] [CrossRef] [Green Version]
  38. Da Rocha-Azevedo, B.; Schmid, S.L. Migrasomes: A new organelle of migrating cells. Cell Res. 2015, 25, 1–2. [Google Scholar] [CrossRef] [Green Version]
  39. Baxter, A.A.; Phan, T.K.; Hanssen, E.; Liem, M.; Hulett, M.D.; Mathivanan, S.; Poon, I.K.H. Analysis of extracellular vesicles generated from monocytes under conditions of lytic cell death. Sci. Rep. 2019, 9, 7538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Tucher, C.; Bode, K.; Schiller, P.; Classen, L.; Birr, C.; Souto-Carneiro, M.M.; Blank, N.; Lorenz, H.M.; Schiller, M. Extracellular Vesicle Subtype Released From Activated or Apoptotic T-Lymphocytes Carry a Specific and Stimulus-Dependent Protein Cargo. Front. Immunol. 2018, 9, 534. [Google Scholar] [CrossRef] [Green Version]
  41. Erozenci, L.A.; Bottger, F.; Bijnsdorp, I.V.; Jimenez, C.R. Urinary exosomal proteins as (pan-)cancer biomarkers: Insights from the proteome. Febs. Lett. 2019, 593, 1580–1597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Colombo, M.; Raposo, G.; Thery, C. Biogenesis, secretion, and intercellular interactions of exosomes and other extracellular vesicles. Annu. Rev. Cell. Dev. Biol. 2014, 30, 255–289. [Google Scholar] [CrossRef] [PubMed]
  43. Nager, A.R.; Goldstein, J.S.; Herranz-Perez, V.; Portran, D.; Ye, F.; Garcia-Verdugo, J.M.; Nachury, M.V. An Actin Network Dispatches Ciliary GPCRs into Extracellular Vesicles to Modulate Signaling. Cell 2017, 168, 252–263.e14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Phua, S.C.; Chiba, S.; Suzuki, M.; Su, E.; Roberson, E.C.; Pusapati, G.V.; Setou, M.; Rohatgi, R.; Reiter, J.F.; Ikegami, K.; et al. Dynamic Remodeling of Membrane Composition Drives Cell Cycle through Primary Cilia Excision. Cell 2017, 168, 264–279.e15. [Google Scholar] [CrossRef] [Green Version]
  45. Park, S.J.; Kim, J.M.; Kim, J.; Hur, J.; Park, S.; Kim, K.; Shin, H.J.; Chwae, Y.J. Molecular mechanisms of biogenesis of apoptotic exosome-like vesicles and their roles as damage-associated molecular patterns. Proc. Natl. Acad. Sci. USA 2018, 115, E11721–E11730. [Google Scholar] [CrossRef] [Green Version]
  46. Zhou, X.X.; Xie, F.; Wang, L.; Zhang, L.; Zhang, S.P.; Fang, M.Y.; Zhou, F.F. The function and clinical application of extracellular vesicles in innate immune regulation. Cell. Mol. Immunol. 2020, 17, 323–334. [Google Scholar] [CrossRef]
  47. Perez-Garijo, A.; Steller, H. Spreading the word: Non-autonomous effects of apoptosis during development, regeneration and disease. Development 2015, 142, 3253–3262. [Google Scholar] [CrossRef] [Green Version]
  48. Sangani, N.B.; Gomes, A.R.; Curfs, L.M.G.; Reutelingsperger, C.P. The role of Extracellular Vesicles during CNS development. Prog. Neurobiol. 2021, 205, 102124. [Google Scholar] [CrossRef]
  49. Simon, C.; Greening, D.W.; Bolumar, D.; Balaguer, N.; Salamonsen, L.A.; Vilella, F. Extracellular Vesicles in Human Reproduction in Health and Disease. Endocr. Rev. 2018, 39, 292–332. [Google Scholar] [CrossRef] [Green Version]
  50. Sproviero, D.; La Salvia, S.; Colombo, F.; Zucca, S.; Pansarasa, O.; Diamanti, L.; Costa, A.; Lova, L.; Giannini, M.; Gagliardi, S.; et al. Leukocyte Derived Microvesicles as Disease Progression Biomarkers in Slow Progressing Amyotrophic Lateral Sclerosis Patients. Front. Neurosci. 2019, 13, 344. [Google Scholar] [CrossRef] [Green Version]
  51. Bobrie, A.; Colombo, M.; Raposo, G.; Thery, C. Exosome Secretion: Molecular Mechanisms and Roles in Immune Responses. Traffic 2011, 12, 1659–1668. [Google Scholar] [CrossRef] [PubMed]
  52. Zakharova, L.; Svetlova, M.; Fomina, A.F. T cell exosomes induce cholesterol accumulation in human monocytes via phosphatidylserine receptor. J. Cell. Physiol. 2007, 212, 174–181. [Google Scholar] [CrossRef]
  53. Muller, G.; Jung, C.; Wied, S.; Biemer-Daub, G.; Frick, W. Transfer of the glycosylphosphatidylinositol-anchored 5′-nucleotidase CD73 from adiposomes into rat adipocytes stimulates lipid synthesis. Brit. J. Pharmacol. 2010, 160, 878–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Thery, C.; Boussac, M.; Veron, P.; Ricciardi-Castagnoli, P.; Raposo, G.; Garin, J.; Amigorena, S. Proteomic analysis of dendritic cell-derived exosomes: A secreted subcellular compartment distinct from apoptotic vesicles. J. Immunol. 2001, 166, 7309–7318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Li, H.; Liao, Y.T.; Gao, L.; Zhuang, T.; Huang, Z.Y.; Zhu, H.M.; Ge, J.B. Coronary Serum Exosomes Derived from Patients with Myocardial Ischemia Regulate Angiogenesis through the miR-939-mediated Nitric Oxide Signaling Pathway. Theranostics 2018, 8, 2079–2093. [Google Scholar] [CrossRef] [PubMed]
  56. Krause, M.; Rak-Raszewska, A.; Naillat, F.; Saarela, U.; Schmidt, C.; Ronkainen, V.P.; Bart, G.; Yla-Herttuala, S.; Vainio, S.J. Exosomes as secondary inductive signals involved in kidney organogenesis. J. Extracell. Vesicles 2018, 7, 1422675. [Google Scholar] [CrossRef] [Green Version]
  57. Deng, F.Y.; Miller, J. A review on protein markers of exosome from different bio-resources and the antibodies used for characterization. J. Histotechnol. 2019, 42, 226–239. [Google Scholar] [CrossRef]
  58. Atkin-Smith, G.K.; Tixeira, R.; Paone, S.; Mathivanan, S.; Collins, C.; Liem, M.; Goodall, K.J.; Ravichandran, K.S.; Hulett, M.D.; Poon, I.K.H. A novel mechanism of generating extracellular vesicles during apoptosis via a beads-on-a-string membrane structure. Nat. Commun. 2015, 6, 8439. [Google Scholar] [CrossRef] [Green Version]
  59. Zhao, X.X.; Lei, Y.X.; Zheng, J.J.; Peng, J.Y.; Li, Y.; Yu, L.; Chen, Y. Identification of markers for migrasome detection. Cell Discov. 2019, 5, 27. [Google Scholar] [CrossRef] [Green Version]
  60. Wu, D.N.; Xu, Y.; Ding, T.L.; Zu, Y.; Yang, C.; Yu, L. Pairing of integrins with ECM proteins determines migrasome formation. Cell Res. 2017, 27, 1397–1400. [Google Scholar] [CrossRef]
  61. Kowal, J.; Tkach, M.; Thery, C. Biogenesis and secretion of exosomes. Curr. Opin. Cell. Biol. 2014, 29, 116–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Zhang, X.; Yuan, X.; Shi, H.; Wu, L.; Qian, H.; Xu, W. Exosomes in cancer: Small particle, big player. J. Hematol. Oncol. 2015, 8, 83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Colombo, M.; Moita, C.; van Niel, G.; Kowal, J.; Vigneron, J.; Benaroch, P.; Manel, N.; Moita, L.F.; Thery, C.; Raposo, G. Analysis of ESCRT functions in exosome biogenesis, composition and secretion highlights the heterogeneity of extracellular vesicles. J. Cell Sci. 2013, 126, 5553–5565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Van Balkom, B.W.; Eisele, A.S.; Pegtel, D.M.; Bervoets, S.; Verhaar, M.C. Quantitative and qualitative analysis of small RNAs in human endothelial cells and exosomes provides insights into localized RNA processing, degradation and sorting. J. Extracell. Vesicles 2015, 4, 26760. [Google Scholar] [CrossRef]
  65. Keerthikumar, S.; Chisanga, D.; Ariyaratne, D.; Saffar, H.; Anand, S.; Zhao, K.N.; Samuel, M.; Pathan, M.; Jois, M.; Chilamkurti, N.; et al. ExoCarta: A Web-Based Compendium of Exosomal Cargo. J. Mol. Biol. 2016, 428, 688–692. [Google Scholar] [CrossRef] [Green Version]
  66. Pathan, M.; Fonseka, P.; Chitti, S.V.; Kang, T.; Sanwlani, R.; Van Deun, J.; Hendrix, A.; Mathivanan, S. Vesiclepedia 2019:a compendium of RNA, proteins, lipids and metabolites in extracellular vesicles. Nucleic Acids Res. 2019, 47, D516–D519. [Google Scholar] [CrossRef] [Green Version]
  67. Ostrowski, M.; Carmo, N.B.; Krumeich, S.; Fanget, I.; Raposo, G.; Savina, A.; Moita, C.F.; Schauer, K.; Hume, A.N.; Freitas, R.P.; et al. Rab27a and Rab27b control different steps of the exosome secretion pathway. Nat. Cell Biol. 2010, 12, 19–30. [Google Scholar] [CrossRef] [Green Version]
  68. Nakase, I.; Ueno, N.; Matsuzawa, M.; Noguchi, K.; Hirano, M.; Omura, M.; Takenaka, T.; Sugiyama, A.; Kobayashi, N.B.; Hashimoto, T.; et al. Environmental pH stress influences cellular secretion and uptake of extracellular vesicles. Febs. Open. Bio. 2021, 11, 753–767. [Google Scholar] [CrossRef]
  69. Messenger, S.W.; Woo, S.S.; Sun, Z.Z.; Martin, T.F.J. A Ca2+-stimulated exosome release pathway in cancer cells is regulated by Munc13-4. J. Cell. Biol. 2019, 218, 1422. [Google Scholar] [CrossRef] [Green Version]
  70. Eitan, E.; Suire, C.; Zhang, S.; Mattson, M.P. Impact of lysosome status on extracellular vesicle content and release. Ageing Res. Rev. 2016, 32, 65–74. [Google Scholar] [CrossRef] [Green Version]
  71. Asare-Werehene, M.; Nakka, K.; Reunov, A.; Chiu, C.T.; Lee, W.T.; Abedini, M.R.; Wang, P.W.; Shieh, D.B.; Dilworth, F.J.; Carmona, E.; et al. The exosome-mediated autocrine and paracrine actions of plasma gelsolin in ovarian cancer chemoresistance. Oncogene 2020, 39, 1600–1616. [Google Scholar] [CrossRef] [Green Version]
  72. Li, I.; Nabet, B.Y. Exosomes in the tumor microenvironment as mediators of cancer therapy resistance. Mol. Cancer 2019, 18, 32. [Google Scholar] [CrossRef]
  73. Segawa, K.; Kurata, S.; Yanagihashi, Y.; Brummelkamp, T.R.; Matsuda, F.; Nagata, S. Caspase-mediated cleavage of phospholipid flippase for apoptotic phosphatidylserine exposure. Science 2014, 344, 1164–1168. [Google Scholar] [CrossRef] [Green Version]
  74. Miyanishi, M.; Tada, K.; Koike, M.; Uchiyama, Y.; Kitamura, T.; Nagata, S. Identification of Tim4 as a phosphatidylserine receptor. Nature 2007, 450, 435–439. [Google Scholar] [CrossRef] [Green Version]
  75. Phuyal, S.; Skotland, T.; Hessvik, N.P.; Simolin, H.; Overbye, A.; Brech, A.; Parton, R.G.; Ekroos, K.; Sandvig, K.; Llorente, A. The Ether Lipid Precursor Hexadecylglycerol Stimulates the Release and Changes the Composition of Exosomes Derived from PC-3 Cells. J. Biol. Chem. 2015, 290, 4225–4237. [Google Scholar] [CrossRef] [Green Version]
  76. Showalter, M.R.; Wancewicz, B.; Fiehn, O.; Archard, J.A.; Clayton, S.; Wagner, J.; Deng, P.; Halmai, J.; Fink, K.D.; Bauer, G.; et al. Primed mesenchymal stem cells package exosomes with metabolites associated with immunomodulation. Biochem. Bioph. Res. Commun. 2019, 512, 729–735. [Google Scholar] [CrossRef]
  77. Wubbolts, R.; Leckie, R.S.; Veenhuizen, P.T.M.; Schwarzmann, G.; Mobius, W.; Hoernschemeyer, J.; Slot, J.W.; Geuze, H.J.; Stoorvogel, W. Proteomic and biochemical analyses of human B cell-derived exosomes—Potential implications for their function and multivesicular body formation. J. Biol. Chem. 2003, 278, 10963–10972. [Google Scholar] [CrossRef] [Green Version]
  78. Carayon, K.; Chaoui, K.; Ronzier, E.; Lazar, I.; Bertrand-Michel, J.; Roques, V.; Balor, S.; Terce, F.; Lopez, A.; Salome, L.; et al. Proteolipidic Composition of Exosomes Changes during Reticulocyte Maturation. J. Biol. Chem. 2011, 286, 34426–34439. [Google Scholar] [CrossRef] [Green Version]
  79. Peterka, O.; Jirasko, R.; Chocholouskova, M.; Kuchar, L.; Wolrab, D.; Hajek, R.; Vrana, D.; Strouhal, O.; Melichar, B.; Holcapek, M. Lipidomic characterization of exosomes isolated from human plasma using various mass spectrometry techniques. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2020, 1865, 158634. [Google Scholar] [CrossRef]
  80. Pols, M.S.; Lumperman, J. Trafficking and function of the tetraspanin CD63. Exp. Cell Res. 2009, 315, 1584–1592. [Google Scholar] [CrossRef]
  81. Mathivanan, S.; Fahner, C.J.; Reid, G.E.; Simpson, R.J. ExoCarta 2012: Database of exosomal proteins, RNA and lipids. Nucleic Acids Res. 2012, 40, D1241–D1244. [Google Scholar] [CrossRef] [Green Version]
  82. Kalluri, R.; LeBleu, V.S. The biology, function, and biomedical applications of exosomes. Science 2020, 367, eaau6977. [Google Scholar] [CrossRef]
  83. Bon, H.; Hales, P.; Lumb, S.; Holdsworth, G.; Johnson, T.; Qureshi, O.; Twomey, B.M. Spontaneous Extracellular Matrix Accumulation in a Human in vitro Model of Renal Fibrosis Is Mediated by α V Integrins. Nephron 2019, 142, 328–350. [Google Scholar] [CrossRef]
  84. Kreidberg, J.A.; Donovan, M.J.; Goldstein, S.L.; Rennke, H.; Shepherd, K.; Jones, R.C.; Jaenisch, R. α 3 β 1 integrin has a crucial role in kidney and lung organogenesis. Development 1996, 122, 3537–3547. [Google Scholar] [CrossRef]
  85. Zhu, H.; Liao, J.; Zhou, X.; Hong, X.; Song, D.; Hou, F.F.; Liu, Y.; Fu, H. Tenascin-C promotes acute kidney injury to chronic kidney disease progression by impairing tubular integrity via alphavbeta6 integrin signaling. Kidney Int. 2020, 97, 1017–1031. [Google Scholar] [CrossRef]
  86. Hayek, S.S.; Leaf, D.E.; Reiser, J. Soluble Urokinase Receptor and Acute Kidney Injury Reply. N. Engl. J. Med. 2020, 382, 2167–2168. [Google Scholar] [CrossRef]
  87. Pang, X.C.; He, X.; Qiu, Z.W.; Zhang, H.X.; Xie, R.; Liu, Z.Y.; Gu, Y.L.; Zhao, N.; Xiang, Q.; Cui, Y.M. Targeting integrin pathways: Mechanisms and advances in therapy. Signal. Transduct. Target. 2023, 8, 1. [Google Scholar] [CrossRef]
  88. Henderson, N.C.; Arnold, T.D.; Katamura, Y.; Giacomini, M.M.; Rodriguez, J.D.; McCarty, J.H.; Pellicoro, A.; Raschperger, E.; Betsholtz, C.; Ruminski, P.G.; et al. Targeting of α(v) integrin identifies a core molecular pathway that regulates fibrosis in several organs. Nat. Med. 2013, 19, 1617–1624. [Google Scholar] [CrossRef] [Green Version]
  89. Liu, B.; Zheng, X.; Meng, F.; Han, Y.; Song, Y.; Liu, F.; Li, S.; Zhang, L.; Gu, F.; Zhang, X.; et al. Overexpression of beta1 integrin contributes to polarity reversal and a poor prognosis of breast invasive micropapillary carcinoma. Oncotarget 2018, 9, 4338–4353. [Google Scholar] [CrossRef] [Green Version]
  90. Wen, S.W.; Lima, L.G.; Lobb, R.J.; Norris, E.L.; Hastie, M.L.; Krumeich, S.; Moller, A. Breast Cancer-Derived Exosomes Reflect the Cell-of-Origin Phenotype. Proteomics 2019, 19, e1800180. [Google Scholar] [CrossRef]
  91. Dhondt, B.; Van Deun, J.; Vermaerke, S.; de Marco, A.; Lumen, N.; De Wever, O.; Hendrix, A. Urinary extracellular vesicle biomarkers in urological cancers: From discovery towards clinical implementation. Int. J. Biochem. Cell Biol. 2018, 99, 236–256. [Google Scholar] [CrossRef]
  92. Huebner, A.R.; Somparn, P.; Benjachat, T.; Leelahavanichkul, A.; Avihingsanon, Y.; Fenton, R.A.; Pisitkun, T. Exosomes in Urine Biomarker Discovery. Urin. Proteom. Kidney Dis. Biomark. Discov. 2015, 845, 43–58. [Google Scholar] [CrossRef]
  93. Faught, E.; Henrickson, L.; Vijayan, M.M. Plasma exosomes are enriched in Hsp70 and modulated by stress and cortisol in rainbow trout. J. Endocrinol. 2017, 232, 237–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Hong, C.S.; Funk, S.; Muller, L.; Boyiadzis, M.; Whiteside, T.L. Isolation of biologically active and morphologically intact exosomes from plasma of patients with cancer. J. Extracell. Vesicles 2016, 5, 55–56. [Google Scholar] [CrossRef] [PubMed]
  95. Zarovni, N.; Corrado, A.; Guazzi, P.; Zocco, D.; Lari, E.; Radano, G.; Muhhina, J.; Fondelli, C.; Gavrilova, J.; Chiesi, A. Integrated isolation and quantitative analysis of exosome shuttled proteins and nucleic acids using immunocapture approaches. Methods 2015, 87, 46–58. [Google Scholar] [CrossRef]
  96. Greening, D.W.; Xu, R.; Ji, H.; Tauro, B.J.; Simpson, R.J. A protocol for exosome isolation and characterization: Evaluation of ultracentrifugation, density-gradient separation, and immunoaffinity capture methods. Methods Mol. Biol. 2015, 1295, 179–209. [Google Scholar] [CrossRef]
  97. Lobb, R.J.; Becker, M.; Wen, S.W.; Wong, C.S.; Wiegmans, A.P.; Leimgruber, A.; Moller, A. Optimized exosome isolation protocol for cell culture supernatant and human plasma. J. Extracell. Vesicles 2015, 4, 27031. [Google Scholar] [CrossRef]
  98. Wang, Z.X.; Wu, H.J.; Fine, D.; Schmulen, J.; Hu, Y.; Godin, B.; Zhang, J.X.J.; Liu, X.W. Ciliated micropillars for the microfluidic-based isolation of nanoscale lipid vesicles. Lab. Chip 2013, 13, 2879–2882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Deregibus, M.C.; Figliolini, F.; D’antico, S.; Manzini, P.M.; Pasquino, C.; De Lena, M.; Tetta, C.; Brizzi, M.F.; Camussi, G. Charge-based precipitation of extracellular vesicles. Int. J. Mol. Med. 2016, 38, 1359–1366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Quintana, J.F.; Makepeace, B.L.; Babayan, S.A.; Ivens, A.; Pfarr, K.M.; Blaxter, M.; Debrah, A.; Wanji, S.; Ngangyung, H.F.; Bah, G.S.; et al. Extracellular Onchocerca-derived small RNAs in host nodules and blood. Parasite Vector 2015, 8, 58. [Google Scholar] [CrossRef] [Green Version]
  101. Xu, R.; Greening, D.W.; Zhu, H.J.; Takahashi, N.; Simpson, R.J. Extracellular vesicle isolation and characterization: Toward clinical application. J. Clin. Investig. 2016, 126, 1152–1162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Momen-Heravi, F.; Balaj, L.; Alian, S.; Trachtenberg, A.J.; Hochberg, F.H.; Skog, J.; Kuo, W.P. Impact of biofluid viscosity on size and sedimentation efficiency of the isolated microvesicles. Front. Physiol. 2012, 3, 162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Cvjetkovic, A.; Lotvall, J.; Lasser, C. The influence of rotor type and centrifugation time on the yield and purity of extracellular vesicles. J. Extracell. Vesicles 2014, 3, 23111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Jeppesen, D.K.; Hvam, M.L.; Primdahl-Bengtson, B.; Boysen, A.T.; Whitehead, B.; Dyrskjot, L.; Orntoft, T.F.; Howard, K.A.; Ostenfeld, M.S. Comparative analysis of discrete exosome fractions obtained by differential centrifugation. J. Extracell. Vesicles 2014, 3, 25011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Takov, K.; Yellon, D.M.; Davidson, S.M. Comparison of small extracellular vesicles isolated from plasma by ultracentrifugation or size-exclusion chromatography: Yield, purity and functional potential. J. Extracell. Vesicles 2019, 8, 1560809. [Google Scholar] [CrossRef] [Green Version]
  106. Tauro, B.J.; Greening, D.W.; Mathias, R.A.; Ji, H.; Mathivanan, S.; Scott, A.M.; Simpson, R.J. Comparison of ultracentrifugation, density gradient separation, and immunoaffinity capture methods for isolating human colon cancer cell line LIM1863-derived exosomes. Methods 2012, 56, 293–304. [Google Scholar] [CrossRef]
  107. Boing, A.N.; van der Pol, E.; Grootemaat, A.E.; Coumans, F.A.; Sturk, A.; Nieuwland, R. Single-step isolation of extracellular vesicles by size-exclusion chromatography. J. Extracell. Vesicles 2014, 3, 23430. [Google Scholar] [CrossRef]
  108. Muller, L.; Hong, C.S.; Stolz, D.B.; Watkins, S.C.; Whiteside, T.L. Isolation of biologically-active exosomes from human plasma. J. Immunol. Methods 2014, 411, 55–65. [Google Scholar] [CrossRef] [Green Version]
  109. Clayton, A.; Court, J.; Navabi, H.; Adams, M.; Mason, M.D.; Hobot, J.A.; Newman, G.R.; Jasani, B. Analysis of antigen presenting cell derived exosomes, based on immuno-magnetic isolation and flow cytometry. J. Immunol. Methods 2001, 247, 163–174. [Google Scholar] [CrossRef]
  110. Rider, M.A.; Hurwitz, S.N.; Meckes, D.G. ExtraPEG: A Polyethylene Glycol-Based Method for Enrichment of Extracellular Vesicles. Sci. Rep. 2016, 6, 23978. [Google Scholar] [CrossRef]
  111. Martinez-Greene, J.A.; Hernandez-Ortega, K.; Quiroz-Baez, R.; Resendis-Antonio, O.; Pichardo-Casas, I.; Sinclair, D.A.; Budnik, B.; Hidalgo-Miranda, A.; Uribe-Querol, E.; Ramos-Godinez, M.D.; et al. Quantitative proteomic analysis of extracellular vesicle subgroups isolated by an optimized method combining polymer-based precipitation and size exclusion chromatography. J. Extracell. Vesicles 2021, 10, 12087. [Google Scholar] [CrossRef] [PubMed]
  112. Caradec, J.; Kharmate, G.; Hosseini-Beheshti, E.; Adomat, H.; Gleave, M.; Guns, E. Reproducibility and efficiency of serum-derived exosome extraction methods. Clin. Biochem. 2014, 47, 1286–1292. [Google Scholar] [CrossRef]
  113. Taylor, D.D.; Zacharias, W.; Gercel-Taylor, C. Exosome isolation for proteomic analyses and RNA profiling. Methods Mol. Biol. 2011, 728, 235–246. [Google Scholar] [CrossRef]
  114. Zlotogorski-Hurvitz, A.; Dayan, D.; Chaushu, G.; Korvala, J.; Salo, T.; Sormunen, R.; Vered, M. Human Saliva-Derived Exosomes: Comparing Methods of Isolation. J. Histochem. Cytochem. 2015, 63, 181–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Weng, Y.; Sui, Z.; Shan, Y.; Hu, Y.; Chen, Y.; Zhang, L.; Zhang, Y. Effective isolation of exosomes with polyethylene glycol from cell culture supernatant for in-depth proteome profiling. Analyst 2016, 141, 4640–4646. [Google Scholar] [CrossRef] [PubMed]
  116. Liga, A.; Vliegenthart, A.D.B.; Oosthuyzen, W.; Dear, J.W.; Kersaudy-Kerhoas, M. Exosome isolation: A microfluidic road-map. Lab. Chip 2015, 15, 2388–2394. [Google Scholar] [CrossRef] [Green Version]
  117. Erdbrugger, U.; Lannigan, J. Analytical Challenges of Extracellular Vesicle Detection: A Comparison of Different Techniques. Cytom. Part A 2016, 89, 123–134. [Google Scholar] [CrossRef]
  118. Sokolova, V.; Ludwig, A.K.; Hornung, S.; Rotan, O.; Horn, P.A.; Epple, M.; Glebel, B. Characterisation of exosomes derived from human cells by nanoparticle tracking analysis and scanning electron microscopy. Colloid. Surface B 2011, 87, 146–150. [Google Scholar] [CrossRef]
  119. Dragovic, R.A.; Gardiner, C.; Brooks, A.S.; Tannetta, D.S.; Ferguson, D.J.P.; Hole, P.; Carr, B.; Redman, C.W.G.; Harris, A.L.; Dobson, P.J.; et al. Sizing and phenotyping of cellular vesicles using Nanoparticle Tracking Analysis. Nanomed.-Nanotechnol. 2011, 7, 780–788. [Google Scholar] [CrossRef] [Green Version]
  120. Petersen, K.E.; Manangon, E.; Hood, J.L.; Wickline, S.A.; Fernandez, D.P.; Johnson, W.P.; Gale, B.K. A review of exosome separation techniques and characterization of B16-F10 mouse melanoma exosomes with AF4-UV-MALS-DLS-TEM. Anal. Bioanal. Chem. 2014, 406, 7855–7866. [Google Scholar] [CrossRef] [Green Version]
  121. Miranda, K.; Girard-Dias, W.; Attias, M.; de Souza, W.; Ramos, I. Three dimensional reconstruction by electron microscopy in the life sciences: An introduction for cell and tissue biologists. Mol. Reprod. Dev. 2015, 82, 530–547. [Google Scholar] [CrossRef]
  122. Cheng, Y.R.; Zeng, Q.Y.; Han, Q.; Xia, W.L. Effect of pH, temperature and freezing-thawing on quantity changes and cellular uptake of exosomes. Protein Cell. 2019, 10, 295–299. [Google Scholar] [CrossRef] [Green Version]
  123. Crescitelli, R.; Lasser, C.; Lotvall, J. Isolation and characterization of extracellular vesicle subpopulations from tissues. Nat. Protoc. 2021, 16, 1548–1580. [Google Scholar] [CrossRef]
  124. Grigoryeva, E.S.; Tashireva, L.A.; Savelieva, O.E.; Zavyalova, M.V.; Popova, N.O.; Kuznetsov, G.A.; Andryuhova, E.S.; Perelmuter, V.M. The Association of Integrins beta3, beta4, and alphaVbeta5 on Exosomes, CTCs and Tumor Cells with Localization of Distant Metastasis in Breast Cancer Patients. Int. J. Mol. Sci. 2023, 24, 2929. [Google Scholar] [CrossRef]
  125. Tayoun, T.; Faugeroux, V.; Oulhen, M.; Aberlenc, A.; Pawlikowska, P.; Farace, F. CTC-Derived Models: A Window into the Seeding Capacity of Circulating Tumor Cells (CTCs). Cells 2019, 8, 1145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Tamkovich, S.; Tutanov, O.; Efimenko, A.; Grigor’eva, A.; Ryabchikova, E.; Kirushina, N.; Vlassov, V.; Tkachuk, V.; Laktionov, P. Blood Circulating Exosomes Contain Distinguishable Fractions of Free and Cell-Surface-Associated Vesicles. Curr. Mol. Med. 2019, 19, 273–285. [Google Scholar] [CrossRef] [PubMed]
  127. Keseru, J.S.; Soltesz, B.; Lukacs, J.; Marton, E.; Szilagyi-Bonizs, M.; Penyige, A.; Poka, R.; Nagy, B. Detection of cell-free, exosomal and whole blood mitochondrial DNA copy number in plasma or whole blood of patients with serous epithelial ovarian cancer. J. Biotechnol. 2019, 298, 76–81. [Google Scholar] [CrossRef] [PubMed]
  128. Mohrmann, L.; Huang, H.J.; Hong, D.S.; Tsimberidou, A.M.; Fu, S.Q.; Piha-Paul, S.A.; Subbiah, V.; Karp, D.D.; Naing, A.; Krug, A.; et al. Liquid Biopsies Using Plasma Exosomal Nucleic Acids and Plasma Cell- Free DNA Compared with Clinical Outcomes of Patients with Advanced Cancers. Clin. Cancer Res. 2018, 24, 181–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Allenson, K.; Castillo, J.; San Lucas, F.A.; Scelo, G.; Kim, D.U.; Bernard, V.; Davis, G.; Kumar, T.; Katz, M.; Overman, M.J.; et al. High prevalence of mutant KRAS in circulating exosome-derived DNA from early-stage pancreatic cancer patients. Ann. Oncol. 2017, 28, 741–747. [Google Scholar] [CrossRef]
  130. Li, M.; Zeringer, E.; Barta, T.; Schageman, J.; Cheng, A.; Vlassov, A.V. Analysis of the RNA content of the exosomes derived from blood serum and urine and its potential as biomarkers. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 2014, 369. [Google Scholar] [CrossRef]
  131. Lamond, A.I. Molecular biology of the cell, 4th edition. Nature 2002, 417, 383. [Google Scholar] [CrossRef]
  132. Camussi, G.; Deregibus, M.C.; Bruno, S.; Cantaluppi, V.; Biancone, L. Exosomes/microvesicles as a mechanism of cell-to-cell communication. Kidney Int. 2010, 78, 838–848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Schneider, A.; Simons, M. Exosomes: Vesicular carriers for intercellular communication in neurodegenerative disorders. Cell Tissue Res. 2013, 352, 33–47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Simons, M.; Raposo, G. Exosomes—Vesicular carriers for intercellular communication. Curr. Opin. Cell Biol. 2009, 21, 575–581. [Google Scholar] [CrossRef]
  135. Webber, J.; Steadman, R.; Mason, M.D.; Tabi, Z.; Clayton, A. Cancer Exosomes Trigger Fibroblast to Myofibroblast Differentiation. Cancer Res. 2010, 70, 9621–9630. [Google Scholar] [CrossRef] [Green Version]
  136. Tian, T.; Wang, Y.; Wang, H.; Zhu, Z.; Xiao, Z. Visualizing of the cellular uptake and intracellular trafficking of exosomes by live-cell microscopy. J. Cell. Biochem. 2010, 111, 488–496. [Google Scholar] [CrossRef] [PubMed]
  137. Torres, V.I.; Barrera, D.P.; Varas-Godoy, M.; Arancibia, D.; Inestrosa, N.C. Selective Surface and Intraluminal Localization of Wnt Ligands on Small Extracellular Vesicles Released by HT-22 Hippocampal Neurons. Front. Cell Dev. Biol. 2021, 9, 735888. [Google Scholar] [CrossRef] [PubMed]
  138. Khalyfa, A.; Gozal, D. Exosomal miRNAs as potential biomarkers of cardiovascular risk in children. J. Transl. Med. 2014, 12, 162. [Google Scholar] [CrossRef] [Green Version]
  139. Jayaseelan, V.P.; Arumugam, P. Exosome-derived ncRNAs as potential drivers of epigenetic reprogramming of cancer stem cells. Epigenomics 2021, 13, 1439–1442. [Google Scholar] [CrossRef]
  140. Shrivastava, S.; Ray, R.M.; Holguin, L.; Echavarria, L.; Grepo, N.; Scott, T.A.; Burnett, J.; Morris, K.V. Exosome-mediated stable epigenetic repression of HIV-1. Nat. Commun. 2021, 12, 5541. [Google Scholar] [CrossRef]
  141. Munich, S.; Sobo-Vujanovic, A.; Buchser, W.J.; Beer-Stolz, D.; Vujanovic, N.L. Dendritic cell exosomes directly kill tumor cells and activate natural killer cells via TNF superfamily ligands. Oncoimmunology 2012, 1, 1074–1083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Guan, S.S.; Li, Q.R.; Liu, P.P.; Xuan, X.Y.; Du, Y. Umbilical cord blood-derived dendritic cells loaded with BGC823 tumor antigens and DC-derived exosomes stimulate efficient cytotoxic T-lymphocyte responses and antitumor immunity in vitro and in vivo. Cent. Eur. J. Immunol. 2014, 39, 142–151. [Google Scholar] [CrossRef] [PubMed]
  143. Mulcahy, L.A.; Pink, R.C.; Carter, D.R. Routes and mechanisms of extracellular vesicle uptake. J. Extracell. Vesicles 2014, 3, 24641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Prada, I.; Meldolesi, J. Binding and Fusion of Extracellular Vesicles to the Plasma Membrane of Their Cell Targets. Int. J. Mol. Sci. 2016, 17, 1296. [Google Scholar] [CrossRef] [Green Version]
  145. Kiss, A.L.; Botos, E. Endocytosis via caveolae: Alternative pathway with distinct cellular compartments to avoid lysosomal degradation? J. Cell. Mol. Med. 2009, 13, 1228–1237. [Google Scholar] [CrossRef] [Green Version]
  146. Feng, D.; Zhao, W.L.; Ye, Y.Y.; Bai, X.C.; Liu, R.Q.; Chang, L.F.; Zhou, Q.; Sui, S.F. Cellular Internalization of Exosomes Occurs Through Phagocytosis. Traffic 2010, 11, 675–687. [Google Scholar] [CrossRef]
  147. Mettlen, M.; Chen, P.-H.; Srinivasan, S.; Danuser, G.; Schmid, S.L. Regulation of Clathrin-Mediated Endocytosis. Annu. Rev. Biochem. 2018, 87, 871–896. [Google Scholar] [CrossRef]
  148. Mercer, J.; Schelhaas, M.; Helenius, A. Virus Entry by Endocytosis. Annu. Rev. Biochem. 2010, 79, 803–833. [Google Scholar] [CrossRef] [Green Version]
  149. Ha, D.; Yang, N.N.; Nadithe, V. Exosomes as therapeutic drug carriers and delivery vehicles across biological membranes: Current perspectives and future challenges. Acta Pharm. Sin. B 2016, 6, 287–296. [Google Scholar] [CrossRef] [Green Version]
  150. Kitai, Y.; Kawasaki, T.; Sueyoshi, T.; Kobiyama, K.; Ishii, K.J.; Zou, J.; Akira, S.; Matsuda, T.; Kawai, T. DNA-Containing Exosomes Derived from Cancer Cells Treated with Topotecan Activate a STING- Dependent Pathway and Reinforce Antitumor Immunity. J. Immunol. 2017, 198, 1649–1659. [Google Scholar] [CrossRef] [Green Version]
  151. Wu, L.; Zhang, X.; Zhang, B.; Shi, H.; Yuan, X.; Sun, Y.; Pan, Z.; Qian, H.; Xu, W. Exosomes derived from gastric cancer cells activate NF-kappaB pathway in macrophages to promote cancer progression. Tumour Biol. 2016, 37, 12169–12180. [Google Scholar] [CrossRef] [PubMed]
  152. Rigalli, J.P.; Barros, E.R.; Sommers, V.; Bindels, R.J.M.; Hoenderop, J.G.J. Novel Aspects of Extracellular Vesicles in the Regulation of Renal Physiological and Pathophysiological Processes. Front. Cell. Dev. Biol. 2020, 8, 244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Van der Lubbe, N.; Jansen, P.M.; Salih, M.; Fenton, R.A.; van den Meiracker, A.H.; Danser, A.H.; Zietse, R.; Hoorn, E.J. The phosphorylated sodium chloride cotransporter in urinary exosomes is superior to prostasin as a marker for aldosteronism. Hypertension 2012, 60, 741–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Ranghino, A.; Dimuccio, V.; Papadimitriou, E.; Bussolati, B. Extracellular vesicles in the urine: Markers and mediators of tissue damage and regeneration. Clin. Kidney J. 2015, 8, 23–30. [Google Scholar] [CrossRef] [PubMed]
  155. Sonoda, H.; Yokota-Ikeda, N.; Oshikawa, S.; Kanno, Y.; Yoshinaga, K.; Uchida, K.; Ueda, Y.; Kimiya, K.; Uezono, S.; Ueda, A.; et al. Decreased abundance of urinary exosomal aquaporin-1 in renal ischemia-reperfusion injury. Am. J. Physiol. Renal. Physiol. 2009, 297, F1006–F1016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Bulacio, R.P.; Anzai, N.; Ouchi, M.; Torres, A.M. Organic Anion Transporter 5 (Oat5) Urinary Excretion Is a Specific Biomarker of Kidney Injury: Evaluation of Urinary Excretion of Exosomal Oat5 after N-Acetylcysteine Prevention of Cisplatin Induced Nephrotoxicity. Chem. Res. Toxicol. 2015, 28, 1595–1602. [Google Scholar] [CrossRef] [Green Version]
  157. Zhou, H.; Cheruvanky, A.; Hu, X.; Matsumoto, T.; Hiramatsu, N.; Cho, M.E.; Berger, A.; Leelahavanichkul, A.; Doi, K.; Chawla, L.S.; et al. Urinary exosomal transcription factors, a new class of biomarkers for renal disease. Kidney Int. 2008, 74, 613–621. [Google Scholar] [CrossRef] [Green Version]
  158. du Cheyron, D.; Daubin, C.; Poggioli, J.; Ramakers, M.; Houillier, P.; Charbonneau, P.; Paillard, M. Urinary measurement of Na+/H+ exchanger isoform 3 (NHE3) protein as new marker of tubule injury in critically ill patients with ARF. Am. J. Kidney Dis. 2003, 42, 497–506. [Google Scholar] [CrossRef]
  159. Sonoda, H.; Lee, B.R.; Park, K.H.; Nihalani, D.; Yoon, J.H.; Ikeda, M.; Kwon, S.H. miRNA profiling of urinary exosomes to assess the progression of acute kidney injury. Sci. Rep. 2019, 9, 4692. [Google Scholar] [CrossRef] [Green Version]
  160. Lv, L.L.; Cao, Y.H.; Ni, H.F.; Xu, M.; Liu, D.; Liu, H.; Chen, P.S.; Liu, B.C. MicroRNA-29c in urinary exosome/microvesicle as a biomarker of renal fibrosis. Am. J. Physiol. Renal. Physiol. 2013, 305, F1220–F1227. [Google Scholar] [CrossRef] [Green Version]
  161. Lv, C.Y.; Zhao, Z.Y.; Yang, T.L.; Wang, Y.L.; Li, B.; Lv, J.; Ding, W.J. Liquid biopsy biomarkers of renal interstitial fibrosis based on urinary exosome. Exp. Mol. Pathol. 2018, 105, 223–228. [Google Scholar] [CrossRef]
  162. Khurana, R.; Ranches, G.; Schafferer, S.; Lukasser, M.; Rudnicki, M.; Mayer, G.; Huttenhofer, A. Identification of urinary exosomal noncoding RNAs as novel biomarkers in chronic kidney disease. RNA 2017, 23, 142–152. [Google Scholar] [CrossRef] [Green Version]
  163. Yu, Y.T.; Bai, F.; Qin, N.; Liu, W.J.; Sun, Q.; Zhou, Y.; Yang, J.W. Non-Proximal Renal Tubule-Derived Urinary Exosomal miR-200b as a Biomarker of Renal Fibrosis. Nephron 2018, 139, 269–282. [Google Scholar] [CrossRef] [PubMed]
  164. Furini, G.; Schroeder, N.; Huang, L.H.; Boocock, D.; Scarpellini, A.; Coveney, C.; Tonoli, E.; Ramaswamy, R.; Ball, G.; Verderio, C.; et al. Proteomic Profiling Reveals the Transglutaminase-2 Externalization Pathway in Kidneys after Unilateral Ureteric Obstruction. J. Am. Soc. Nephrol. 2018, 29, 880–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Cao, Y.H.; Shi, Y.H.; Wang, Y.W.; Yang, Y.L.; Guo, W.J.; Zhang, C.F.; Pei, W.J.; Fu, C. Exosomal hsa_circ_0008925 from Urine Is Related to Chronic Renal Fibrosis. Dis. Markers 2022, 2022, 1899282. [Google Scholar] [CrossRef] [PubMed]
  166. Cao, Y.H.; Shi, Y.H.; Yang, Y.L.; Wu, Z.L.; Peng, N.N.; Xiao, J.; Dou, F.; Xu, J.J.; Pei, W.J.; Fu, C.; et al. Urinary exosomes derived circRNAs as biomarkers for chronic renal fibrosis. Ann. Med. 2022, 54, 1966–1976. [Google Scholar] [CrossRef]
  167. Zubiri, I.; Posada-Ayala, M.; Sanz-Maroto, A.; Calvo, E.; Martin-Lorenzo, M.; Gonzalez-Calero, L.; de la Cuesta, F.; Lopez, J.A.; Fernandez-Fernandez, B.; Ortiz, A.; et al. Diabetic nephropathy induces changes in the proteome of human urinary exosomes as revealed by label-free comparative analysis. J. Proteom. 2014, 96, 92–102. [Google Scholar] [CrossRef]
  168. Rossi, L.; Nicoletti, M.C.; Carmosino, M.; Mastrofrancesco, L.; Di Franco, A.; Indrio, F.; Lella, R.; Laviola, L.; Giorgino, F.; Svelto, M.; et al. Urinary Excretion of Kidney Aquaporins as Possible Diagnostic Biomarker of Diabetic Nephropathy. J. Diabetes Res. 2017, 2017, 4360357. [Google Scholar] [CrossRef] [Green Version]
  169. Delic, D.; Eisele, C.; Schmid, R.; Baum, P.; Wiech, F.; Gerl, M.; Zimdahl, H.; Pullen, S.S.; Urquhart, R. Urinary Exosomal miRNA Signature in Type II Diabetic Nephropathy Patients. PLoS ONE 2016, 11, e0150154. [Google Scholar] [CrossRef]
  170. Eissa, S.; Matboli, M.; Aboushahba, R.; Bekhet, M.M.; Soliman, Y. Urinary exosomal microRNA panel unravels novel biomarkers for diagnosis of type 2 diabetic kidney disease. J. Diabetes Complicat. 2016, 30, 1585–1592. [Google Scholar] [CrossRef]
  171. Eissa, S.; Matboli, M.; Bekhet, M.M. Clinical verification of a novel urinary microRNA panal: 133b,-342 and-30 as biomarkers for diabetic nephropathy identified by bioinformatics analysis. Biomed. Pharmacother. 2016, 83, 92–99. [Google Scholar] [CrossRef]
  172. Li, W.T.; Yang, S.; Qiao, R.; Zhang, J. Potential Value of Urinary Exosome-Derived let-7c-5p in the Diagnosis and Progression of Type II Diabetic Nephropathy. Clin. Lab. 2018, 64, 709–718. [Google Scholar] [CrossRef]
  173. Hogan, M.C.; Manganelli, L.; Woollard, J.R.; Masyuk, A.I.; Masyuk, T.V.; Tammachote, R.; Huang, B.Q.; Leontovich, A.A.; Beito, T.G.; Madden, B.J.; et al. Characterization of PKD Protein-Positive Exosome-Like Vesicles. J. Am. Soc. Nephrol. 2009, 20, 278–288. [Google Scholar] [CrossRef] [Green Version]
  174. Hogan, M.C.; Bakeberg, J.L.; Gainullin, V.G.; Irazabal, M.V.; Harmon, A.J.; Lieske, J.C.; Charlesworth, M.C.; Johnson, K.L.; Madden, B.J.; Zenka, R.M.; et al. Identification of Biomarkers for PKD1 Using Urinary Exosomes. J. Am. Soc. Nephrol. 2015, 26, 1661–1670. [Google Scholar] [CrossRef] [Green Version]
  175. Salih, M.; Demmers, J.A.; Berstarosti, K.; Leonhard, W.N.; Losekoot, M.; van Kooten, C.; Gansevoort, R.T.; Peters, D.J.M.; Zietse, R.; Hoorn, E.J.; et al. Proteomics of Urinary Vesicles Links Plakins and Complement to Polycystic Kidney Disease. J. Am. Soc. Nephrol. 2016, 27, 3079–3092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Keri, K.C.; Regner, K.R.; Dall, A.T.; Park, F. Urinary exosomal expression of activator of G protein signaling 3 in polycystic kidney disease. BMC Res. Notes 2018, 11, 359. [Google Scholar] [CrossRef] [PubMed]
  177. Bruschi, M.; Granata, S.; Santucci, L.; Candiano, G.; Fabris, A.; Antonucci, N.; Petretto, A.; Bartolucci, M.; Del Zotto, G.; Antonini, F.; et al. Proteomic Analysis of Urinary Microvesicles and Exosomes in Medullary Sponge Kidney Disease and Autosomal Dominant Polycystic Kidney Disease. Clin. J. Am. Soc. Nephro. 2019, 14, 834–843. [Google Scholar] [CrossRef] [PubMed]
  178. Raimondo, F.; Morosi, L.; Corbetta, S.; Chinello, C.; Brambilla, P.; Della Mina, P.; Villa, A.; Albo, G.; Battaglia, C.; Bosari, S.; et al. Differential protein profiling of renal cell carcinoma urinary exosomes. Mol. Biosyst. 2013, 9, 1220–1233. [Google Scholar] [CrossRef] [Green Version]
  179. Butz, H.; Nofech-Mozes, R.; Ding, Q.; Khella, H.W.Z.; Szabo, P.M.; Jewett, M.; Finelli, A.; Lee, J.; Ordon, M.; Stewart, R.; et al. Exosomal MicroRNAs Are Diagnostic Biomarkers and Can Mediate Cell-Cell Communication in Renal Cell Carcinoma. Eur. Urol. Focus. 2016, 2, 210–218. [Google Scholar] [CrossRef]
  180. Song, S.Q.; Long, M.M.; Yu, G.P.; Cheng, Y.J.; Yang, Q.; Liu, J.Y.; Wang, Y.W.; Sheng, J.Y.; Wang, L.H.; Wang, Z.; et al. Urinary exosome miR-30c-5p as a biomarker of clear cell renal cell carcinoma that inhibits progression by targeting HSPA5. J. Cell. Mol. Med. 2019, 23, 6755–6765. [Google Scholar] [CrossRef] [Green Version]
  181. Kurahashi, R.; Kadomatsu, T.; Baba, M.; Hara, C.; Itoh, H.; Miyata, K.; Endo, M.; Morinaga, J.; Terada, K.; Araki, K.; et al. MicroRNA-204-5p: A novel candidate urinary biomarker of Xp11.2 translocation renal cell carcinoma. Cancer Sci. 2019, 110, 1897–1908. [Google Scholar] [CrossRef] [Green Version]
  182. Zarbock, A.; Nadim, M.K.; Pickkers, P.; Gomez, H.; Bell, S.; Joannidis, M.; Kashani, K.; Koyner, J.L.; Pannu, N.; Meersch, M.; et al. Sepsis-associated acute kidney injury: Consensus report of the 28th Acute Disease Quality Initiative workgroup. Nat. Rev. Nephrol. 2023, 19, 401–417. [Google Scholar] [CrossRef] [PubMed]
  183. Zhou, H.; Pisitkun, T.; Aponte, A.; Yuen, P.S.; Hoffert, J.D.; Yasuda, H.; Hu, X.; Chawla, L.; Shen, R.F.; Knepper, M.A.; et al. Exosomal Fetuin-A identified by proteomics: A novel urinary biomarker for detecting acute kidney injury. Kidney Int. 2006, 70, 1847–1857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Da-Silva, C.C.S.; Anauate, A.C.; Guirao, T.P.; Novaes, A.D.; Maquigussa, E.; Boim, M.A. Analysis of exosome-derived microRNAs as early biomarkers of lipopolysaccharide-induced acute kidney injury in rats. Front. Physiol. 2022, 13, 944864. [Google Scholar] [CrossRef]
  185. Awdishu, L.; Tsunoda, S.; Pearlman, M.; Kokoy-Mondragon, C.; Ghassemian, M.; Naviaux, R.K.; Patton, H.M.; Mehta, R.L.; Vijay, B.; RamachandraRao, S.P. Identification of Maltase Glucoamylase as a Biomarker of Acute Kidney Injury in Patients with Cirrhosis. Crit. Care Res. Pract. 2019, 2019, 5912804. [Google Scholar] [CrossRef] [Green Version]
  186. Ferguson, M.A.; Vaidya, V.S.; Waikar, S.S.; Collings, F.B.; Sunderland, K.E.; Gioules, C.J.; Bonventre, J.V. Urinary liver-type fatty acid-binding protein predicts adverse outcomes in acute kidney injury. Kidney Int. 2010, 77, 708–714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Mishra, J.; Dent, C.; Tarabishi, R.; Mitsnefes, M.M.; Ma, Q.; Kelly, C.; Ruff, S.M.; Zahedi, K.; Shao, M.; Bean, J.; et al. Neutrophil gelatinase-associated lipocalin (NGAL) as a biomarker for acute renal injury after cardiac surgery. Lancet 2005, 365, 1231–1238. [Google Scholar] [CrossRef]
  188. Singer, E.; Elger, A.; Elitok, S.; Kettritz, R.; Nickolas, T.L.; Barasch, J.; Luft, F.C.; Schmidt-Ott, K.M. Urinary neutrophil gelatinase-associated lipocalin distinguishes pre-renal from intrinsic renal failure and predicts outcomes. Kidney Int. 2011, 80, 405–414. [Google Scholar] [CrossRef] [Green Version]
  189. Han, W.K.; Bailly, V.; Abichandani, R.; Thadhani, R.; Bonventre, J.V. Kidney Injury Molecule-1 (KIM-1): A novel biomarker for human renal proximal tubule injury. Kidney Int. 2002, 62, 237–244. [Google Scholar] [CrossRef] [Green Version]
  190. Sheira, G.; Noreldin, N.; Tamer, A.; Saad, M. Urinary biomarker N-acetyl-β-D-glucosaminidase can predict severity of renal damage in diabetic nephropathy. J. Diabetes Metab. Disord. 2015, 14, 4. [Google Scholar] [CrossRef] [Green Version]
  191. Meersch, M.; Schmidt, C.; Van Aken, H.; Martens, S.; Rossaint, J.; Singbartl, K.; Gorlich, D.; Kellum, J.A.; Zarbock, A. Urinary TIMP-2 and IGFBP7 as Early Biomarkers of Acute Kidney Injury and Renal Recovery following Cardiac Surgery. PLoS ONE 2014, 9, e93460. [Google Scholar] [CrossRef] [Green Version]
  192. Wetz, A.J.; Richardt, E.M.; Wand, S.; Kunze, N.; Schotola, H.; Quintel, M.; Brauer, A.; Moerer, O. Quantification of urinary TIMP-2 and IGFBP-7: An adequate diagnostic test to predict acute kidney injury after cardiac surgery? Crit. Care 2015, 19, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Kasuno, K.; Shirakawa, K.; Yoshida, H.; Mori, K.; Kimura, H.; Takahashi, N.; Nobukawa, Y.; Shigemi, K.; Tanabe, S.; Yamada, N.; et al. Renal redox dysregulation in AKI: Application for oxidative stress marker of AKI. Am. J. Physiol. Renal. Physiol. 2014, 307, F1342–F1351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Kashani, K.; Al-Khafaji, A.; Ardiles, T.; Artigas, A.; Bagshaw, S.M.; Bell, M.; Bihorac, A.; Birkhahn, R.; Cely, C.M.; Chawla, L.S.; et al. Discovery and validation of cell cycle arrest biomarkers in human acute kidney injury. Crit. Care 2013, 17, R25. [Google Scholar] [CrossRef] [Green Version]
  195. Yokoi, S.; Kasuno, K.; Nishimori, K.; Nishikawa, S.; Nishikawa, Y.; Morita, S.; Kobayashi, M.; Fukushima, S.; Mikami, D.; Takahashi, N.; et al. Analytical and clinical validation of rapid chemiluminescence enzyme immunoassay for urinary thioredoxin, an oxidative stress-dependent early biomarker of acute kidney injury. Clin. Chim. Acta 2020, 507, 271–279. [Google Scholar] [CrossRef]
  196. Global, K.D.I. Kidney Disease: Improving Global Outcomes (KDIGO) CKD-MBD Update Work Group. KDIGO 2017 Clinical Practice Guideline Update for the Diagnosis, Evaluation, Prevention, and Treatment of Chronic Kidney Disease-Mineral and Bone Disorder (CKD-MBD). Kidney Int. Suppl. 2017, 7, E1. [Google Scholar] [CrossRef] [Green Version]
  197. Lange, T.; Artelt, N.; Kindt, F.; Stracke, S.; Rettig, R.; Lendeckel, U.; Chadjichristos, C.E.; Kavvadas, P.; Chatziantoniou, C.; Endlich, K.; et al. MiR-21 is up-regulated in urinary exosomes of chronic kidney disease patients and after glomerular injury. J. Cell. Mol. Med. 2019, 23, 4839–4843. [Google Scholar] [CrossRef] [Green Version]
  198. Gudehithlu, K.P.; Hart, P.; Joshi, A.; Garcia-Gomez, I.; Cimbaluk, D.J.; Dunea, G.; Arruda, J.A.L.; Singh, A.K. Urine exosomal ceruloplasmin: A potential early biomarker of underlying kidney disease. Clin. Exp. Nephrol. 2019, 23, 1013–1021. [Google Scholar] [CrossRef]
  199. Wang, S.X.; Zhang, J.J.; Nauli, S.M.; Li, X.G.; Starremans, P.G.; Luo, Y.; Roberts, K.A.; Zhou, J. Fibrocystin/polyductin, found in the same protein complex with polycystin-2, regulates calcium responses in kidney epithelia. Mol. Cell. Biol. 2007, 27, 3241–3252. [Google Scholar] [CrossRef] [Green Version]
  200. Fan, P.D.; Jiang, Z.; Diamond, I.; Yao, L.N. Up-Regulation of AGS3 during Morphine Withdrawal Promotes cAMP Superactivation via Adenylyl Cyclase 5 and 7 in Rat Nucleus Accumbens/Striatal Neurons. Mol. Pharmacol. 2009, 76, 526–533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Nadella, R.; Blumer, J.B.; Jia, G.F.; Kwon, M.; Akbulut, T.; Qian, F.; Sedlic, F.; Wakatsuki, T.; Sweeney, W.E.; Wilson, P.D.; et al. Activator of G Protein Signaling 3 Promotes Epithelial Cell Proliferation in PKD. J. Am. Soc. Nephrol. 2010, 21, 1275–1280. [Google Scholar] [CrossRef] [Green Version]
  202. Regner, K.R.; Nozu, K.; Lanier, S.M.; Blumer, J.B.; Avner, E.D.; Sweeney, W.E.; Park, F. Loss of activator of G-protein signaling 3 impairs renal tubular regeneration following acute kidney injury in rodents. Faseb. J. 2011, 25, 1844–1855. [Google Scholar] [CrossRef] [Green Version]
  203. Patschan, D.; Buschmann, I.; Ritter, O.; Kribben, A. Cell-Based Therapies in Acute Kidney Injury (AKI). Kidney Blood Press. Res. 2018, 43, 673–681. [Google Scholar] [CrossRef]
  204. Xiong, G.; Tang, W.; Zhang, D.; He, D.; Wei, G.; Atala, A.; Liang, X.J.; Bleyer, A.J.; Bleyer, M.E.; Yu, J.; et al. Impaired Regeneration Potential in Urinary Stem Cells Diagnosed from the Patients with Diabetic Nephropathy. Theranostics 2019, 9, 4221–4232. [Google Scholar] [CrossRef]
  205. Yun, C.W.; Lee, S.H. Potential and Therapeutic Efficacy of Cell-based Therapy Using Mesenchymal Stem Cells for Acute/chronic Kidney Disease. Int. J. Mol. Sci. 2019, 20, 1619. [Google Scholar] [CrossRef] [Green Version]
  206. Ohno, S.; Takanashi, M.; Sudo, K.; Ueda, S.; Ishikawa, A.; Matsuyama, N.; Fujita, K.; Mizutani, T.; Ohgi, T.; Ochiya, T.; et al. Systemically Injected Exosomes Targeted to EGFR Deliver Antitumor MicroRNA to Breast Cancer Cells. Mol. Ther. 2013, 21, 185–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Alvarez-Erviti, L.; Seow, Y.Q.; Yin, H.F.; Betts, C.; Lakhal, S.; Wood, M.J.A. Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes. Nat. Biotechnol. 2011, 29, 341–345. [Google Scholar] [CrossRef] [PubMed]
  208. Wahlgren, J.; Karlson, T.D.; Brisslert, M.; Sani, F.V.; Telemo, E.; Sunnerhagen, P.; Valadi, H. Plasma exosomes can deliver exogenous short interfering RNA to monocytes and lymphocytes. Nucleic Acids Res. 2012, 40, e130. [Google Scholar] [CrossRef] [Green Version]
  209. EL Andaloussi, S.; Lakhal, S.; Mager, I.; Wood, M.J.A. Exosomes for targeted siRNA delivery across biological barriers. Adv. Drug. Deliv. Rev. 2013, 65, 391–397. [Google Scholar] [CrossRef] [PubMed]
  210. Izco, M.; Blesa, J.; Schleef, M.; Schmeer, M.; Porcari, R.; Al-Shawi, R.; Ellmerich, S.; de Toro, M.; Gardiner, C.; Seow, Y.; et al. Systemic Exosomal Delivery of shRNA Minicircles Prevents Parkinsonian Pathology. Mol. Ther. 2019, 27, 2111–2122. [Google Scholar] [CrossRef] [PubMed]
  211. Sun, D.M.; Zhuang, X.Y.; Xiang, X.Y.; Liu, Y.L.; Zhang, S.Y.; Liu, C.R.; Barnes, S.; Grizzle, W.; Miller, D.; Zhang, H.G. A Novel Nanoparticle Drug Delivery System: The Anti-inflammatory Activity of Curcumin Is Enhanced When Encapsulated in Exosomes. Mol. Ther. 2010, 18, 1606–1614. [Google Scholar] [CrossRef]
  212. Zhuang, X.Y.; Xiang, X.Y.; Grizzle, W.; Sun, D.M.; Zhang, S.Q.; Axtell, R.C.; Ju, S.W.; Mu, J.Y.; Zhang, L.F.; Steinman, L.; et al. Treatment of Brain Inflammatory Diseases by Delivering Exosome Encapsulated Anti-Inflammatory Drugs From the Nasal Region to the Brain. Mol. Ther. 2012, 20, 239. [Google Scholar] [CrossRef] [PubMed]
  213. Yang, T.Z.; Martin, P.; Fogarty, B.; Brown, A.; Schurman, K.; Phipps, R.; Yin, V.P.; Lockman, P.; Bai, S.H. Exosome Delivered Anticancer Drugs Across the Blood-Brain Barrier for Brain Cancer Therapy in Danio Rerio. Pharm. Res. 2015, 32, 2003–2014. [Google Scholar] [CrossRef] [Green Version]
  214. Tian, Y.H.; Li, S.P.; Song, J.; Ji, T.J.; Zhu, M.T.; Anderson, G.J.; Wei, J.Y.; Nie, G.J. A doxorubicin delivery platform using engineered natural membrane vesicle exosomes for targeted tumor therapy. Biomaterials 2014, 35, 2383–2390. [Google Scholar] [CrossRef]
  215. Tacar, O.; Sriamornsak, P.; Dass, C.R. Doxorubicin: An update on anticancer molecular action, toxicity and novel drug delivery systems. J. Pharm. Pharmacol. 2013, 65, 157–170. [Google Scholar] [CrossRef] [PubMed]
  216. Pascucci, L.; Cocce, V.; Bonomi, A.; Ami, D.; Ceccarelli, P.; Ciusani, E.; Vigano, L.; Locatelli, A.; Sisto, F.; Doglia, S.M.; et al. Paclitaxel is incorporated by mesenchymal stromal cells and released in exosomes that inhibit in vitro tumor growth: A new approach for drug delivery. J. Control. Release 2014, 192, 262–270. [Google Scholar] [CrossRef] [PubMed]
  217. Jang, S.C.; Kim, O.Y.; Yoon, C.M.; Choi, D.S.; Roh, T.Y.; Park, J.; Nilsson, J.; Lotvall, J.; Kim, Y.K.; Gho, Y.S. Bioinspired Exosome-Mimetic Nanovesicles for Targeted Delivery of Chemotherapeutics to Malignant Tumors. ACS Nano 2013, 7, 7698–7710. [Google Scholar] [CrossRef] [PubMed]
  218. Lee, J.; Kim, J.; Jeong, M.; Lee, H.; Goh, U.; Kim, H.; Kim, B.; Park, J.H. Liposome-Based Engineering of Cells To Package Hydrophobic Compounds in Membrane Vesicles for Tumor Penetration. Nano Lett. 2015, 15, 2938–2944. [Google Scholar] [CrossRef] [PubMed]
  219. Haney, M.J.; Zhao, Y.L.; Harrison, E.B.; Mahajan, V.; Ahmed, S.; He, Z.J.; Suresh, P.; Hingtgen, S.D.; Klyachko, N.L.; Mosley, R.L.; et al. Specific Transfection of Inflamed Brain by Macrophages: A New Therapeutic Strategy for Neurodegenerative Diseases. PLoS ONE 2013, 8, e61852. [Google Scholar] [CrossRef] [Green Version]
  220. Maguire, C.A.; Balaj, L.; Sivaraman, S.; Crommentuijn, M.H.W.; Ericsson, M.; Mincheva-Nilsson, L.; Baranov, V.; Gianni, D.; Tannous, B.A.; Sena-Esteves, M.; et al. Microvesicle-associated AAV Vector as a Novel Gene Delivery System. Mol. Ther. 2012, 20, 960–971. [Google Scholar] [CrossRef] [Green Version]
  221. Wang, X.Y.; Zhang, H.Y.; Yang, H.O.; Bai, M.; Ning, T.; Li, S.; Li, J.L.; Deng, T.; Ying, G.G.; Ba, Y. Cell-derived Exosomes as Promising Carriers for Drug Delivery and Targeted Therapy. Curr. Cancer Drug. Tar. 2018, 18, 347–354. [Google Scholar] [CrossRef] [PubMed]
  222. Haney, M.J.; Klyachko, N.L.; Zhaoa, Y.L.; Gupta, R.; Plotnikova, E.G.; He, Z.J.; Patel, T.; Piroyan, A.; Sokolsky, M.; Kabanov, A.V.; et al. Exosomes as drug delivery vehicles for Parkinson’s disease therapy. J. Control. Release 2015, 207, 18–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Pomatto, M.A.C.; Bussolati, B.; D’Antico, S.; Ghiotto, S.; Tetta, C.; Brizzi, M.F.; Camussi, G. Improved Loading of Plasma-Derived Extracellular Vesicles to Encapsulate Antitumor miRNAs. Mol. Ther. Meth. Clin. Dev. 2019, 13, 133–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Yang, Z.G.; Shi, J.F.; Xie, J.; Wang, Y.F.; Sun, J.Y.; Liu, T.Z.; Zhao, Y.R.; Zhao, X.T.; Wang, X.M.; Ma, Y.F.; et al. Large-scale generation of functional mRNA-encapsulating exosomes via cellular nanoporation. Nat. Biomed. Eng. 2020, 4, 69–83. [Google Scholar] [CrossRef]
  225. Fuhrmann, G.; Serio, A.; Mazo, M.; Nair, R.; Stevens, M.M. Active loading into extracellular vesicles significantly improves the cellular uptake and photodynamic effect of porphyrins. J. Control. Release 2015, 205, 35–44. [Google Scholar] [CrossRef]
  226. Tang, T.T.; Wang, B.; Wu, M.; Li, Z.L.; Feng, Y.; Cao, J.Y.; Yin, D.; Liu, H.; Tang, R.N.; Crowley, S.D.; et al. Extracellular vesicle-encapsulated IL-10 as novel nanotherapeutics against ischemic AKI. Sci. Adv. 2020, 6, eaaz0748. [Google Scholar] [CrossRef]
  227. Patel, D.M.; Shah, J.; Srivastava, A.S. Therapeutic Potential of Mesenchymal Stem Cells in Regenerative Medicine. Stem Cells Int. 2013, 2013, 496218. [Google Scholar] [CrossRef] [Green Version]
  228. Wu, X.M.; Jiang, J.; Gu, Z.K.; Zhang, J.Y.; Chen, Y.; Liu, X.L. Mesenchymal stromal cell therapies: Immunomodulatory properties and clinical progress. Stem Cell. Res. Ther. 2020, 11, 345. [Google Scholar] [CrossRef]
  229. Gatti, S.; Bruno, S.; Deregibus, M.C.; Sordi, A.; Cantaluppi, V.; Tetta, C.; Camussi, G. Microvesicles derived from human adult mesenchymal stem cells protect against ischaemia-reperfusion-induced acute and chronic kidney injury. Nephrol. Dial. Transpl. 2011, 26, 1474–1483. [Google Scholar] [CrossRef] [Green Version]
  230. Biancone, L.; Bruno, S.; Deregibus, M.C.; Tetta, C.; Camussi, G. Therapeutic potential of mesenchymal stem cell-derived microvesicles. Nephrol. Dial. Transpl. 2012, 27, 3037–3042. [Google Scholar] [CrossRef]
  231. Bagno, L.; Hatzistergos, K.E.; Balkan, W.; Hare, J.M. Mesenchymal Stem Cell-Based Therapy for Cardiovascular Disease: Progress and Challenges. Mol. Ther. 2018, 26, 1610–1623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Liew, L.C.; Katsuda, T.; Gailhouste, L.; Nakagama, H.; Ochiya, T. Mesenchymal stem cell-derived extracellular vesicles: A glimmer of hope in treating Alzheimer’s disease. Int. Immunol. 2017, 29, 11–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Wang, B.; Yao, K.; Huuskes, B.M.; Shen, H.H.; Zhuang, J.L.; Godson, C.; Brennan, E.P.; Wilkinson-Berka, J.L.; Wise, A.F.; Ricardo, S.D. Mesenchymal Stem Cells Deliver Exogenous MicroRNA-let7c via Exosomes to Attenuate Renal Fibrosis. Mol. Ther. 2016, 24, 1290–1301. [Google Scholar] [CrossRef] [Green Version]
  234. Alzahrani, F.A. Melatonin improves therapeutic potential of mesenchymal stem cells-derived exosomes against renal ischemia-reperfusion injury in rats. Am. J. Transl. Res. 2019, 11, 2887–2907. [Google Scholar]
  235. Yoon, Y.M.; Lee, J.H.; Song, K.H.; Noh, H.; Lee, S.H. Melatonin-stimulated exosomes enhance the regenerative potential of chronic kidney disease-derived mesenchymal stem/stromal cells via cellular prion proteins. J. Pineal Res. 2020, 68, e12632. [Google Scholar] [CrossRef]
  236. Tang, T.T.; Lv, L.L.; Wang, B.; Cao, J.Y.; Feng, Y.; Li, Z.L.; Wu, M.; Wang, F.M.; Wen, Y.; Zhou, L.T.; et al. Employing Macrophage-Derived Microvesicle for Kidney-Targeted Delivery of Dexamethasone: An Efficient Therapeutic Strategy against Renal Inflammation and Fibrosis. Theranostics 2019, 9, 4740–4755. [Google Scholar] [CrossRef]
  237. Zhang, C.Y.; Shang, Y.N.; Chen, X.N.; Midgley, A.D.; Wang, Z.Y.; Zhu, D.S.; Wu, J.; Chen, P.; Wu, L.L.; Wang, X.; et al. Supramolecular Nanofibers Containing Arginine-Glycine-Aspartate (RGD) Peptides Boost Therapeutic Efficacy of Extracellular Vesicles in Kidney Repair. ACS Nano 2020, 14, 12133–12147. [Google Scholar] [CrossRef] [PubMed]
  238. Jo, W.; Kim, J.; Yoon, J.; Jeong, D.; Cho, S.; Jeong, H.; Yoon, Y.J.; Kim, S.C.; Gho, Y.S.; Park, J. Large-scale generation of cell-derived nanovesicles. Nanoscale 2014, 6, 12056–12064. [Google Scholar] [CrossRef] [Green Version]
  239. Yang, Q.; Wang, Y.X.; Pei, G.C.; Deng, X.; Jiang, H.Y.; Wu, J.L.; Zhou, C.; Guo, Y.; Yao, Y.; Zeng, R.; et al. Bone marrow-derived Ly6C(-) macrophages promote ischemia-induced chronic kidney disease. Cell Death Dis. 2019, 10. [Google Scholar] [CrossRef] [Green Version]
  240. Wang, N.; Liang, H.W.; Zen, K. Molecular mechanisms that influence the macrophage M1-M2 polarization balance. Front. Immunol. 2014, 5, 614. [Google Scholar] [CrossRef] [Green Version]
  241. Xie, X.C.; Yang, X.; Wu, J.X.; Tang, S.J.; Yang, L.L.; Fei, X.; Wang, M. Exosome from indoleamine 2,3-dioxygenase-overexpressing bone marrow mesenchymal stem cells accelerates repair process of ischemia/reperfusion-induced acute kidney injury by regulating macrophages polarization. Stem Cell. Res. Ther. 2022, 13, 367. [Google Scholar] [CrossRef] [PubMed]
  242. Shen, B.; Liu, J.; Zhang, F.; Wang, Y.; Qin, Y.; Zhou, Z.H.; Qiu, J.X.; Fan, Y. CCR2 Positive Exosome Released by Mesenchymal Stem Cells Suppresses Macrophage Functions and Alleviates Ischemia/Reperfusion-Induced Renal Injury. Stem Cells Int. 2016, 2016, 1240301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Jin, J.; Shi, Y.F.; Gong, J.G.; Zhao, L.; Li, Y.W.; He, Q.; Huang, H. Exosome secreted from adipose-derived stem cells attenuates diabetic nephropathy by promoting autophagy flux and inhibiting apoptosis in podocyte. Stem Cell. Res. Ther. 2019, 10, 95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Zou, X.Y.; Zhang, G.Y.; Cheng, Z.L.; Yin, D.M.; Du, T.; Ju, G.Q.; Miao, S.; Liu, G.H.; Lu, M.J.; Zhu, Y.J. Microvesicles derived from human Wharton’s Jelly mesenchymal stromal cells ameliorate renal ischemia-reperfusion injury in rats by suppressing CX3CL1. Stem Cell. Res. Ther. 2014, 5, 40. [Google Scholar] [CrossRef] [Green Version]
  245. Gu, D.; Zou, X.Y.; Ju, G.Q.; Zhang, G.Y.; Bao, E.D.; Zhu, Y.J. Mesenchymal Stromal Cells Derived Extracellular Vesicles Ameliorate Acute Renal Ischemia Reperfusion Injury by Inhibition of Mitochondrial Fission through miR-30. Stem Cells Int. 2016, 2016, 2093940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Cao, H.; Cheng, Y.; Gao, H.; Zhuang, J.; Zhang, W.; Bian, Q.; Wang, F.; Du, Y.; Li, Z.; Kong, D.; et al. In Vivo Tracking of Mesenchymal Stem Cell-Derived Extracellular Vesicles Improving Mitochondrial Function in Renal Ischemia-Reperfusion Injury. ACS Nano 2020, 14, 4014–4026. [Google Scholar] [CrossRef]
  247. Li, X.R.; Liao, J.; Su, X.J.; Li, W.Q.; Bi, Z.R.; Wang, J.L.; Su, Q.; Huang, H.T.; Wei, Y.C.; Gao, Y.F.; et al. Human urine-derived stem cells protect against renal ischemia/reperfusion injury in a rat model via exosomal miR-146a-5p which targets IRAK1. Theranostics 2020, 10, 9561–9578. [Google Scholar] [CrossRef]
  248. Zhang, Y.M.; Wang, J.X.; Yang, B.X.; Qiao, R.; Li, A.W.; Guo, H.; Ding, J.; Li, H.; Ye, H.; Wu, D.; et al. Transfer of MicroRNA-216a-5p From Exosomes Secreted by Human Urine-Derived Stem Cells Reduces Renal Ischemia/Reperfusion Injury. Front. Cell. Dev. Biol. 2020, 8, 610587. [Google Scholar] [CrossRef]
  249. Ti, D.D.; Hao, H.J.; Tong, C.A.; Liu, J.J.; Dong, L.; Zheng, J.X.; Zhao, Y.L.; Liu, H.L.; Fu, X.B.; Han, W.D. LPS-preconditioned mesenchymal stromal cells modify macrophage polarization for resolution of chronic inflammation via exosome-shuttled let-7b. J. Transl. Med. 2015, 13, 308. [Google Scholar] [CrossRef] [Green Version]
  250. Song, N.N.; Zhang, T.; Xu, X.L.; Lu, Z.H.; Yu, X.F.; Fang, Y.; Hu, J.C.; Jia, P.; Teng, J.; Ding, X.Q. miR-21 Protects Against Ischemia/Reperfusion-Induced Acute Kidney Injury by Preventing Epithelial Cell Apoptosis and Inhibiting Dendritic Cell Maturation. Front. Physiol. 2018, 9, 790. [Google Scholar] [CrossRef] [Green Version]
  251. Lv, L.L.; Feng, Y.; Wu, M.; Wang, B.; Li, Z.L.; Zhong, X.; Wu, W.J.; Chen, J.; Ni, H.F.; Tang, T.T.; et al. Exosomal miRNA-19b-3p of tubular epithelial cells promotes M1 macrophage activation in kidney injury. Cell Death Differ. 2020, 27, 210–226. [Google Scholar] [CrossRef] [PubMed]
  252. Zhang, W.; Zheng, X.X.; Yu, Y.S.; Zheng, L.; Lan, J.H.; Wu, Y.; Liu, H.; Zhao, A.; Huang, H.; Chen, W. Renal cell carcinoma-derived exosomes deliver lncARSR to induce macrophage polarization and promote tumor progression via STAT3 pathway. Int. J. Biol. Sci. 2022, 18, 3209–3222. [Google Scholar] [CrossRef] [PubMed]
  253. Qu, L.; Ding, J.; Chen, C.; Wu, Z.J.; Liu, B.; Gao, Y.; Chen, W.; Liu, F.; Sun, W.; Li, X.F.; et al. Exosome-Transmitted lncARSR Promotes Sunitinib Resistance in Renal Cancer by Acting as a Competing Endogenous RNA. Cancer Cell. 2016, 29, 653–668. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Extracellular vesicle classification. A depiction of the four different classes of extracellular vesicles indicating their modes of biogenesis and release. Exosomes are generated through the endocytosis of multivesicular bodies and are released via exocytosis, are spherical in shape, and vary in size. Microvesicles are generated and released through budding/shedding from plasma membrane, are irregular in shape, and vary in size. Apoptotic bodies are released through blebbing by cells undergoing apoptosis. Migrasomes (pomegranate-like structures) grow on and are released from the tips or intersections of retraction fibers, which mark the path of migrating cells.
Figure 1. Extracellular vesicle classification. A depiction of the four different classes of extracellular vesicles indicating their modes of biogenesis and release. Exosomes are generated through the endocytosis of multivesicular bodies and are released via exocytosis, are spherical in shape, and vary in size. Microvesicles are generated and released through budding/shedding from plasma membrane, are irregular in shape, and vary in size. Apoptotic bodies are released through blebbing by cells undergoing apoptosis. Migrasomes (pomegranate-like structures) grow on and are released from the tips or intersections of retraction fibers, which mark the path of migrating cells.
Genes 14 01367 g001
Figure 2. Exosome biogenesis. Exosomes are generated in a process that involves the double invagination of the plasma membrane in the early endosomes, resulting in the formation of intracellular multivesicular bodies (MVBs) that possess cell surface proteins and soluble proteins associated with the extracellular environment and the plasma membrane of the parent cell. Within these MVBs, inward invaginations occur, resulting in the formation of intraluminal vesicles (ILVs) in late endosomes following cargo sorting. Both ESCRT-dependent and ESCRT-independent driven pathways participate in creating multivesicular bodies. Exocytic MVBs fuse with the plasma membrane in Rab-GTPases-regulated manner. ILVs eventually become exosomes when secreted to the extracellular microenvironment. Exosome content depends on the cell type and the physiological and pathological condition of the cell.
Figure 2. Exosome biogenesis. Exosomes are generated in a process that involves the double invagination of the plasma membrane in the early endosomes, resulting in the formation of intracellular multivesicular bodies (MVBs) that possess cell surface proteins and soluble proteins associated with the extracellular environment and the plasma membrane of the parent cell. Within these MVBs, inward invaginations occur, resulting in the formation of intraluminal vesicles (ILVs) in late endosomes following cargo sorting. Both ESCRT-dependent and ESCRT-independent driven pathways participate in creating multivesicular bodies. Exocytic MVBs fuse with the plasma membrane in Rab-GTPases-regulated manner. ILVs eventually become exosomes when secreted to the extracellular microenvironment. Exosome content depends on the cell type and the physiological and pathological condition of the cell.
Genes 14 01367 g002
Figure 3. Hallmarks of exosomes. Exosomes contain a wide variety of molecules, such as nucleic acids, proteins, and lipids. The contents of the exosome are a representation of the cell of origin and of the physiological state of the cell from which the exosome is released from.
Figure 3. Hallmarks of exosomes. Exosomes contain a wide variety of molecules, such as nucleic acids, proteins, and lipids. The contents of the exosome are a representation of the cell of origin and of the physiological state of the cell from which the exosome is released from.
Genes 14 01367 g003
Figure 4. Exosome–recipient cell communication. Exosomes communicate with and modulate recipient cell signaling via multiple routes. Upon reaching the target cell, (1) ligands on the exosome surface membrane and receptors on the plasma membrane of the target cell can interact, inducing downstream signaling cascade in the recipient cell. (2) Exosome membrane can fuse with the plasma membrane and release its contents directly into the cytosol of the recipient cell. (3) Exosomes are internalized by the recipient cell and follow the endosomal pathway for cargo release/recycling.
Figure 4. Exosome–recipient cell communication. Exosomes communicate with and modulate recipient cell signaling via multiple routes. Upon reaching the target cell, (1) ligands on the exosome surface membrane and receptors on the plasma membrane of the target cell can interact, inducing downstream signaling cascade in the recipient cell. (2) Exosome membrane can fuse with the plasma membrane and release its contents directly into the cytosol of the recipient cell. (3) Exosomes are internalized by the recipient cell and follow the endosomal pathway for cargo release/recycling.
Genes 14 01367 g004
Table 1. Characterization of extracellular vesicles.
Table 1. Characterization of extracellular vesicles.
FeatureExosomesMicrovesiclesApoptotic BodiesMigrasomes
Size~30–150 nm100–1000 nm50–5000 nm500–3000 nm
MorphologySpherical or cup-shapedHeterogenousHeterogenousHeterogenous
BiogenesisInward budding of MVBs formed in late endosomeBlebbing or outward budding of the plasma membraneRepeated blebbing and retraction of plasma membrane of apoptotic cells Grow on the tips of retraction fibers trailing behind migrating cells
Release Fusion of MVBs with the plasma membrane (exocytosis)Pinching/shedding of the plasma membraneBlebbing of the plasma membrane Breaking of the retraction fibers
Density1.13–1.19 g/mL [51,52]1.25–1.30 g/mL [53]1.16–1.28 g/mL N/A
Composition Proteins, lipids, DNA, and different RNA speciesProteins, lipids, DNA, and different RNA speciesHistones, cellular organelles, and fragmented DNA [54]Contractile, cytoskeleton, cell adhesion and RNA-binding proteins, enzymes, and chaperons [37]
MarkersTetraspanins (CD9, CD63, CD81), TSG101, Alix, flottilin-1 [35,55,56,57]Integrins, selectins, CD40, flotillin-2Annexin V (PS positive) [58]TSPAN4, NDST1, PIGK, EOGT, CPQ [37,59,60]
Abbreviations used: MVB: multivesicular bodies, ALIX: apoptosis-linked gene-2-interacting protein X, TSG101: tumor susceptibility gene 101, TSPAN4: tetraspanin-4, PS: phosphatidylserine, NDST1: bifunctionalheparan sulfate N-deacetylase/N-sulfotransferase 1, PIGK: phosphatidylinositol glycan anchor biosynthesis class K, CPQ: carboxypeptidase Q and EOGT: EGF domain-specific O-linked N-acetylglucosaminetransferase.
Table 2. Methods for exosome isolation and characterization.
Table 2. Methods for exosome isolation and characterization.
Separation TechnologySpecific ApproachAdvantagesLimitations
Centrifugation-based technologyUltracentrifugation
[103,104,105]
Obtains highly purified exosome fractions
-
Low yield, labor-intensive
-
Sediments other vesicles and protein aggregates
-
Good for exosomes purified from cell culture medium, but not bodily fluids
Density gradient centrifugation
[96,106]
Pure preparations
-
Loss of sample
-
Time-consuming
Size-based technologySize exclusion chromatography
[105,107,108]
Preserves the integrity and activity of the exosome, no risk of vesicle aggregation and protein complex formation
-
Run times are long
-
Breaks up larger vesicles, potentially skewing the results
UltrafiltrationSimple protocol, exosome yield has uniform size, high protein and RNA yield
-
Low purity of exosomes
Exosome capture-based technologyImmune affinity capture/magnetic beads
[35,109]
Highly specific and collects exosomes with high purity
-
Low exosome yield
-
Not appropriate for purification of substantial amounts of exosomes
Polymer precipitation-based technologyUse of a solution containing polyethylene glycol, commercial kit most commonly used is ExoQuick
[110,111,112,113,114,115]
Enriches exosomes from large volumes, less labor-intensive, permits reliable and high-throughput isolation of exosomes from low sample volumes
-
Co-isolation of protein aggregates and non-vesicular contaminants such as lipo-proteins
Microfluidic-based technologyImmune-affinity, sieving and trapping on porous structures
[116]
Extremely sensitive and quantitative analysis of exosomes
-
Low exosome yield and presence of exosome aggregation
Therefore, the method of choice would depend on the sample source/type (cell culture media or bodily fluids) and the intended use for the exosomes. Moreover, the different isolation techniques can affect the analysis; as such, making a choice for the appropriate technique should be approached with caution [117].
Table 3. Summary of the urine-derived exosomal biomarkers in kidney diseases.
Table 3. Summary of the urine-derived exosomal biomarkers in kidney diseases.
Kidney DiseaseSource of ExosomeConclusionsReferences
AKIRat urineDecreased exosomal AQP-1 in animals with renal IR.[155]
Human urineElevated exosomal Oat5 in cisplatin-induced AKI.[156]
Rat urineElevated exosomal Fetuin-A in cisplatin-induced AKI.[157]
Elevated exosomal Na+/H+ exchange type-3 in acute tubular necrosis.[158]
Human urineElevated exosomal NGAL and ATF3 in sepsis-induced AKI patients.[157]
Rat urineElevated exosomal levels of miR-16, miR-24, and miR-200c at an early phase of renal IR; elevated exosomal miR-125 and miR-351 at a late phase of renal IR.[159]
CKDHuman urine Decreased exosomal miR-29c associated with degree of renal fibrosis.[160,161]
Mouse kidneyDecreased exosomal miR-181a.[162]
Human urineDecreased exosomal miR-200b was in CKD patients. The decrease was highest in exosomes derived from non-proximal tubule renal cells.[163]
Mouse kidneyIncreased level of secreting transglutaminase-2 in UUO mice.[164]
Increased exosomal expression level of hsa_circ_0008925 in glomerular disease.[165]
Human urineIncreased exosomal expression level of has_circ_0036649 in glomerular disease.[166]
Human urineElevated exosomal bikunin precursor and histone-lysine N-methyltransferase but decreased VDAC1 in diabetic nephropathy patients.[167]
Human urineIncreased levels of AQP-2 and AQP-5 were detected in exosomes derived from diabetic nephropathy patients.[168]
Human urineIncreased levels of microRNAs such as miR-371b-5p, miR-320c, miR-572, miR-1234-5p, miR-6068, miR-6133, miR-4270, miR-4739, and miR-638 derived from exosomes in type 2 diabetic nephropathy patients.[169]
Human urineDecreased levels of miR-30d-5p and miR-30e-5p in type 2 diabetic nephropathy patients.[169]
Human urineElevated levels of miR-15b, miR-30a, miR-34a, miR-133b, miR-342, and miR-636 in exosomes from type 2 diabetic nephropathy patients.[170,171]
Human urineElevated exosomal levels of let-7c-5p but decreased levels of miR-29c-5p and miR-15b-5p in type 2 DN patients.[172]
PKDHuman urineDecreased levels of PC-1 and PC-2 but increased level of TMEM2 in exosomes derived from ADPKD patients with PKD1 mutation.[15,173,174]
Human urineIncreased expression of cystin and ADP-ribosylation factor-like 6 in PKD patients.[173]
Human urineIncreased levels of complement C3 and C9 in urinary EVs derived from ADPKD patients with or without progressive CKD; however, envoplakin, periplakin, and villin-1 levels were only increased in exosomes from ADPKD patients with progressive CKD.[175]
Human and rat urineIncreased exosomal level of AGS3 in PKD animals and patients.[176]
Human urineIncreased exosomal level of prominin 1 (CD133), cellular repressor of E1A-stimulated genes 1 (CREG1), and cadherin 4 in ADPKD patients.[177]
RCCHuman urineIncreased levels of MMP-9, CP, PODXL, DKK4, and CAIX, and decreased levels of AQP1, CD10, DPEP 1, EMMPRIN, and syntenin-1 in the urinary exosomes of RCC patients.Increased levels of CP and PODXL could be used to distinguish RCC patients from healthy control individuals.[178]
Human urineIncreased level of miR-150-5p and decreased level of miR-126-3p have been reported. In addition, exosomal miR-126-3p combined with miR-449a could discriminate healthy individuals from ccRCC patients. [179]
Human urineExosomal miR-30c-5p and miR-204-5p could serve as potential diagnostic biomarkers for early-stage ccRCC[180,181]
Abbreviations used: AQP—aquaporin; Oat5—organic anion transporter 5; Na+/H+—sodium/hydrogen; NGAL—neutrophil gelatinase-associated lipocalin; ATF3—activating transcription factor 3; UUO—unilateral ureteral obstruction; VDAC1—voltage-dependent anion-selective channel protein 1; DN—diabetic nephropathy; PC-1—polycystin-1; PC-2 —polycystin-2; TMEM2– transmembrane protein 2; AGS3—neutrophil gelatinase-associated lipocalin; CD133—prominin 1, CREG1—cellular repressor of E1A-stimulated genes 1; MMP-9—matrix metalloproteinase-9; CP—ceruloplasmin; PODXL—podocalyxin; DKK4—Dickkopf-related protein 4; CAIX—carbonic anhydrase IX; AQP1—aquaporin-1; CD10—neprilysin; DPEP 1—dipeptidase 1; EMMPRIN—extracellular matrix metalloproteinase inducer; ccRCC—clear cell renal cell carcinoma.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Agborbesong, E.; Bissler, J.; Li, X. Liquid Biopsy at the Frontier of Kidney Diseases: Application of Exosomes in Diagnostics and Therapeutics. Genes 2023, 14, 1367. https://doi.org/10.3390/genes14071367

AMA Style

Agborbesong E, Bissler J, Li X. Liquid Biopsy at the Frontier of Kidney Diseases: Application of Exosomes in Diagnostics and Therapeutics. Genes. 2023; 14(7):1367. https://doi.org/10.3390/genes14071367

Chicago/Turabian Style

Agborbesong, Ewud, John Bissler, and Xiaogang Li. 2023. "Liquid Biopsy at the Frontier of Kidney Diseases: Application of Exosomes in Diagnostics and Therapeutics" Genes 14, no. 7: 1367. https://doi.org/10.3390/genes14071367

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop