Next Article in Journal
Ursodeoxycholic Acid Binds PERK and Ameliorates Neurite Atrophy in a Cellular Model of GM2 Gangliosidosis
Next Article in Special Issue
Extracellular Vesicles as Markers of Liver Function: Optimized Workflow for Biomarker Identification in Liver Disease
Previous Article in Journal
A Mixture Method for Robust Detection HCV Early Diagnosis Biomarker with ML Approach and Molecular Docking
Previous Article in Special Issue
Intravesicular Genomic DNA Enriched by Size Exclusion Chromatography Can Enhance Lung Cancer Oncogene Mutation Detection Sensitivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extracellular Vesicles in Breast Cancer: From Biology and Function to Clinical Diagnosis and Therapeutic Management

1
INSERM U538, CRSA, Saint-Antoine University Hospital, 75012 Paris, France
2
INTEGRACELL SAS, 91160 Longjumeau, France
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(8), 7208; https://doi.org/10.3390/ijms24087208
Submission received: 8 March 2023 / Revised: 3 April 2023 / Accepted: 9 April 2023 / Published: 13 April 2023
(This article belongs to the Special Issue The Role of Exosomes in Cancer Diagnosis and Therapy)

Abstract

:
Breast cancer (BC) is the first worldwide most frequent cancer in both sexes and the most commonly diagnosed in females. Although BC mortality has been thoroughly declining over the past decades, there are still considerable differences between women diagnosed with early BC and when metastatic BC is diagnosed. BC treatment choice is widely dependent on precise histological and molecular characterization. However, recurrence or distant metastasis still occurs even with the most recent efficient therapies. Thus, a better understanding of the different factors underlying tumor escape is mainly mandatory. Among the leading candidates is the continuous interplay between tumor cells and their microenvironment, where extracellular vesicles play a significant role. Among extracellular vesicles, smaller ones, also called exosomes, can carry biomolecules, such as lipids, proteins, and nucleic acids, and generate signal transmission through an intercellular transfer of their content. This mechanism allows tumor cells to recruit and modify the adjacent and systemic microenvironment to support further invasion and dissemination. By reciprocity, stromal cells can also use exosomes to profoundly modify tumor cell behavior. This review intends to cover the most recent literature on the role of extracellular vesicle production in normal and cancerous breast tissues. Specific attention is paid to the use of extracellular vesicles for early BC diagnosis, follow-up, and prognosis because exosomes are actually under the spotlight of researchers as a high-potential source of liquid biopsies. Extracellular vesicles in BC treatment as new targets for therapy or efficient nanovectors to drive drug delivery are also summarized.

1. Introduction

Breast cancer (BC) is the first worldwide most commonly diagnosed cancer for both combined sexes and accounts for 11.7% of total cancer incidence and 6.9% of cancer-related deaths. In 2020, more than 2.3 million new BC cases were reported worldwide, with almost 685,000 related deaths, according to WHO. The death rate was considerably higher in developing versus developed countries (15.0 vs. 12.8 per 100,000) [1]. In Europe and the USA, approximately 523,000 and 276,500 new BC cases are diagnosed annually [2], while 138,000 and 42,000 die of BC-related diseases. Although BC mortality has been thoroughly declining over the last decades, there are still large differences between women diagnosed with early BC (considered curable with ∼96% 5-year survival probability in Europe) or when metastatic BC is diagnosed (mostly not curable with ∼38% 5-year survival rate) [3]. The two significant supports of BC management are locoregional treatment and systemic therapy, and both histological and molecular characteristics of BC broadly impact treatment choices.
Nevertheless, even with the development of new efficient therapies (immunotherapy with PDL1 inhibitors), BC recurrence and metastasis are still the leading causes of death [4], mainly because of the emergence of therapy-resistant cancer cells [5,6]. Thus, answering essential questions involving the factors or mechanisms determining the distant metastasis or the acquisition of therapy resistance is strongly mandatory to develop novel effective therapeutic strategies against BC [7]. Recent studies have shown the importance of tumor evolution in the continuous interplay between tumor cells and their microenvironment (cancer cells’ interaction with themselves, bidirectional interaction of cancer cells with stromal cells, etc.) [8,9]. Such communication strategies require specific mechanisms, including direct cell-to-cell contacts but also autocrine, juxtacrine, paracrine, and even endocrine secretion of specific factors (growth factors, matrixins, cytokines, and chemokines) [10]. Among such secreted means are figure exosomes [11], a generic consensus term used to describe any small lipid bilayer-delimited particles that are unable to replicate and are extracellularly released by every cell (including prokaryotic ones) [12,13]. Exosome surface receptors allow their targeting and capture by a broad range of recipient cells that will incorporate either proteic, lipidic, or genetic messages resulting in modifications of their physiological behavior. These exosomes have recently been proven to be efficient means of communication in human diseases [14], especially in cancer [15,16]. As the field of exosomes is highly active [17,18,19], we aimed to review the respective roles of cancer cell-derived exosomes as well as stromal-derived exosomes in BC to better understand the cellular and molecular mechanisms underlying their generation and development. We also emphasize exosomes as powerful tools to efficiently diagnose, better stage and improve BC prognosis, and better design, in a personalized approach, treatment strategies.

2. Extracellular Vesicle Nature, Structure, and Properties

Various amounts of 40–1000 nm membrane-derived extracellular vesicles, continuously released from the plasma membrane (plasma membrane) into the local environment by either eukaryotic or prokaryotic cells, can be detected in almost all biological fluids. The 2018 MISEV consensus rule recommends differentiating label bilayered membrane vesicles such as SEVs (smaller than 200 nm, small extracellular vesicles or exosomes) and MLEVs (larger than 200 nm, medium large EVs or ectosomes) shed by live cells [20,21] from apoptotic bodies or necrotic blebs of a plasma membrane that are the consequences of dying cells disassembly [22].

2.1. Extracellular Vesicle Biogenesis

While MLEVs are heterogeneous membranous vesicles generally deriving from outward plasma membrane budding (ectosome release) [23], SEVs originate from the endosomal compartment [24].

2.1.1. SEVs Biogenesis

SEVs production follows complicated sorting routes and requires several complex protein systems [25,26,27]. SEVs biogenesis starts with the inward budding of small portions of the plasma membrane containing an outer membrane exposed material. These small intracellular vesicles form the early endosome, which subsequently matures and transforms into a late endosome. Inward budding of the limiting membrane of the late endosome then occurs, resulting in the progressive assemblage of intraluminal bilayered vesicles (ILVs) within so-called large multivesicular endosomes or multivesicular bodies (MVBs) (Figure 1). At this step, it seems that the endocytosed cargo is first delivered to the trans-Golgi network (cargo-in) and then back transferred to MVB (cargo-out) [28]. This route seems to follow the MVB-driven Golgi protein quality control pathway that will further degrade miss-folded proteins in endolysosomes [29]. During this process, cytosolic proteins, as well as nucleic acids, can be trapped into ILVs through the action of the ESCRT (endosomal sorting complex required for transport) machinery [30,31] or by following a ceramide ESCRT-independent pathway, suggesting a critical role for lipid raft microdomains in MVBs formation [32]. ESCRT is a family of proteins that associate in successive complexes (ESCRT-0, -I, -II, and -III) at MVBs membrane to sort ubiquitinated cargos into late endosomes [33].
The fate of MVBs varies according to the proteins that are expressed on their surface that will be specifically recognized on the target membrane [34]. Acidification along the endocytic pathway also seems to be required for the degradation and recycling of internalized components [35]. Intracellular calcium [36] and local hypoxia [37] also seem to be major determinants for MVBs degradation versus secretion. Most MVBs are directed for cargo degradation into lysosomes by fusing with them [38,39]). It was demonstrated that the autophagosome may fuse with MVB in a pre-lysosomal step, resulting in a hybrid organelle called amphisome [40], suggesting that both the autophagy degradation process and the exosomal release are closely linked [41]. As MVBs also contain intraluminal proteins and lipids, which are not intended for lysosome degradation, ILVs can release their content into the cytoplasm by direct back-fusion with the endosome-limiting membrane [42]. However, a subset of them fuse to the plasma membrane and release their content externally in the form of SEVs, a secretion process called exosome biogenesis [43,44].
MVBs that are fated for exocytosis are transported to the plasma membrane along cytoskeletal structures such as actin or microtubules and their associated molecular motors kinesins and dyneins [45,46]. MVBs docking to the plasma membrane is strongly regulated by the Rab family of small GTPases proteins [47], especially Rab27a which mediates the docking, tethering, and fusion of MVBs with the plasma membrane [48,49]. Once docked, secretory MVBs couple to the SNARE (soluble N-ethylmaleimide-sensitive component attachment protein receptor) membrane fusion machinery [50]. SNARE complex formation and membrane fusion are tightly controlled by multiple regulatory mechanisms [51], which include the phosphorylation profile of SNARE proteins, which influence either SNARE complex localization or interaction with SNARE partners [52]. SNARE assembly and disassembly mediate the fusion of MVBs with the cell membrane, thus releasing outside the cell the MVBs particles [53] that become exosomes.

2.1.2. MLEVs Biogenesis

Interestingly, in contrast with the importance of MVBs in the SEV formation pathway, MLEV release is totally MVB-independent and does not require exocytosis [54,55]. MLEV are assembled by the regulated outward budding of plasma membrane domains [56], a mechanism depending either on caveolae [57] or clathrin-coated vesicles [58], explaining why ectosomes are surrounded by phospholipid membranes containing lipid rafts and caveolae [59].
However, despite the distinct mechanism for biogenesis and membrane origin, both endosome-origin SEVs and MLEVs can work similarly, and the crucial difference between them has not yet been elucidated [23].

2.2. Extracellular Vesicle Composition

MLEVs, which bud directly from the plasma membrane of healthy cells, contain cytoskeleton and endoplasmic reticulum elements [60,61]. It is considered that MLEVs’ composition mainly reflects the surface proteins of parental cells [23]. As phosphatidylserine repositioning within the cell plasma membrane is a critical factor in MLEVs evagination [62], MLEVs display high levels of phosphatidylserine. In contrast, SEVs have lower ones exposed to the outer membrane leaflet [63].
Because of their endosomal origin, and since they derived from the ILVs in MVBs, SEVs biogenesis heavily depends on the mechanisms that regulate MVBs maturation and trafficking. Along the different sorting mechanisms needed for SEVs production, specific molecules (proteins, lipids, amino acids, metabolites, nucleic acids such as nuclear or mitochondrial DNA or several RNA species, etc.) [64,65] are incorporated into SEVs generating cargo diversity [66,67,68].
SEVs mostly contain proteins originating from the cytosol, the endosomal compartment, and the plasma membrane [12]. Cytosolic protein engulfment involves proteins close to the MVB outer membrane during its inward budding. Proteins such as Heat shock proteins (Hsp90, Hsc70, Hsp60, Hsp20, Hsp27, etc.), growth factors and cytokines (TNF-α, TGF-β, TRAIL, etc.), and enzymes (belonging to central metabolisms such as glycolysis, citric acid cycle, etc.) can be found in exosome lumen [69]. As the budding and release of SEVs require inner plasma membrane actin polymerization, and then the actomyosin cytoskeleton contraction, cytoskeleton proteins such as actin, actinin, dynamin, myosins, and tubulin are also generally found in SEVs [45]. It holds the same for essential regulators of extracellular vesicles trafficking: ESCRT complex proteins and important ESCRT partners molecules implicated in ESCRT assembly or nucleation such as ALIX [70], members of the Rab family [71], and SNARE membrane fusion machinery, required explicitly for MVBs docking and fusion with plasma membrane [72], are also found in SEVs (Figure 2). Tetraspanins (mainly CD9, CD63, CD37, CD81, CD82, and CD53), which are highly conserved integral membrane proteins displaying a high affinity for cholesterol and sphingolipids such as ceramides, are involved in ESCRT-independent exosome release [73] and greatly influence exosome biogenesis and composition [74,75]. They play essential roles in plasma membrane protein scaffolding and anchoring in cellular membranes by creating specific plasma membrane tetraspanin-enriched microdomains [76]), thus facilitating their sorting into SEVs [77,78]. Thus, antigen-presenting molecules (Major Histocompatibility complex MHC class 1 and MHC class 2), glycoproteins (O-linked and N-linked glycans), adhesion molecules (integrins, selectins, etc.), and signaling receptors (TNF receptor, Transferrin receptor, etc.) are frequently found on SEVs membrane [79].
SEVs can also comprise nucleic acid molecules [80]. Various RNA species (mRNAs, rRNA, tRNA, snRNA, snoRNA, piRNA, Y-RNA, scRNA, SRP-RNA, 7SK-RNA, miRNAs (miRs), lncRNAs, circRNAs, etc.) can be evidenced in SEVs [81,82]. Numerous reports have shown the ability of SEVs RNAs to profoundly impact the functional properties of cells that incorporate them [81]. Nuclear and mitochondrial DNA molecules can also be conveyed by SEVs [83,84]. In plasma, cell-free DNA (CFDNA) circulates in both free form and enclosed in SEVs [83]. While large intact DNA is generally associated with MLEVs [85] and is mainly attached to the outer surface of extracellular vesicles [86], shorter double-strand 150 to 6000 bp fragments resulting from DNA fragmentation by DNAses are usually found in SEVs [87]. As CFDNA sometimes harbors mutations, it may reflect the mutational status of parental DNA [88] and serve as a relevant tumor biological marker in liquid biopsies [89,90]. Aside from this complex protein and nucleic acid repertoire in SEVs, metabolomic studies reveal that SEVs contain different classes of low-molecular-weight compounds.
Organic acids, nucleotides, sugars, their derivatives, carnitines, vitamins, related metabolites, and amines are frequently evidenced in SEVs [91]. Lipids (phosphatidylserine, cholesterol, sphingomyelins, and ceramide) participating in intercellular signaling and also ensuring structural stability are present [92]. These metabolites may originate from specific sorting but are more probably synthesized in situ in SEVs as complete, but more often, partial metabolic routes can be evidenced [93].
Extracellular vesicles, released in body fluids, vastly differ in size, content, morphology, and biological mechanisms [94,95]. A single cell line can continuously generate morphologically diverse vesicles [96]. However, little is known about the essential mechanisms that may account for the combinatorial repertoires of SEV cargo and the heterogeneity in cargo compositions across different extracellular vesicle populations and subtypes [97]. Many reports using either proteomics, transcriptomics, or metabolomics have shown how the vesicular protein cargo is distinct from its original sample [98,99,100]. Increasing evidence has pointed to the selectivity in cargo loading during SEV biogenesis rather than a generic regulation of cargo sorting into SEVs [101]. Under hypoxic conditions, tumor cells show changes in morphology, distribution, and accumulation cargo of MVBs. These modifications are associated with significant differences in the number, morphology, and cargo of SEVs [34]. It was also reported that polymorphonuclear neutrophil cells could produce a broad spectrum of SEVs, depending on the environmental conditions prevailing during SEV genesis [102]. Consequently, SEV composition does not simply represent parental cell protein composition, but more notably, SEVs cargo multifaceted variety reflects a specific signature of these source cells at a definite time [103,104].
SEV biogenesis threshold will vary significantly between cell types according to their physiological/pathological status. The high rate of SEVs secretion found in transformed cells suggests that the balance between MVB degradation and secretion is disrupted in cancer toward SEVs cargo release [105]. Such modification is not specific to cancer cells but may also occur in non-transformed ones. In antigen-presenting cells, large amounts of SEVs are found to be released upon stimulation [106].

2.3. Extracellular Vesicle Fate

Once released, SEVs circulate locoregionally or distantly to deliver their cargo content to recipient cells. The encapsulated cargo of SEVs is protected from degradation [79]. Circulating labeled SEVs’ half-life has been evaluated in mice to be about 2 minutes, but detecting SEVs in the bloodstream remains possible several hours after injection [107]. SEVs then use their lipid membranes to enter recipient cells to release cargo. Although a non-specific uptake is shared by every cell type [108], specific targeting to recipient cells is generally required to deliver exosome cargo and exert its function [109]. When reaching the target cell, SEVs can either trigger signaling by directly interacting with extracellular receptors or be uptaken by direct fusion with the plasma membrane or get internalized. Most reports indicate that endocytosis generally internalizes SEVs into the endosomal compartment [110], while the exact mechanisms underlying SEV endocytosis processes remain unclear [111]. Various other mechanisms have also been proposed, including clathrin-mediated endocytosis, caveolin-dependent endocytosis, lipid raft-dependent endocytosis, micropinocytosis, and phagocytosis [112]. After internalization, SEVs can interact within the recipient cell, inducing intracellular signaling and changes to molecular processes that may affect various functions such as apoptosis, autophagy, growth, cell cycle, migration, invasion, and differentiation [113,114].

3. Extracellular Vesicle’s Role in Normal Breast Tissue

3.1. Extracellular Vesicles Production in Normal Mammary Tissue

The mammary gland is one of the very few organs in which substantial development occurs only after birth, undergoing cycles of growth, differentiation, milk secretion, apoptosis, regression, and remodeling during the lifetime of the organism [115]. It develops predominantly during the postnatal period from several invading cells derived from the ectoderm [116]. These cells undergo a morphogenetic program that leads to the development of a series of branching ducts that terminate in sac-like lobules embedded in stromal tissue. Both secretory acini and ducts are lined by an epithelium [117] that later expands to generate a complicated network to deliver milk to newborn progeny. This continuous epithelium consists of an outer basal layer of contractile myoepithelial cells and an inner layer of luminal cells surrounding the lumen. The epithelium includes a subset of stem cells closely interacting with the environment to drive their fate and the ultimate mammary gland phenotype [118,119]. The surrounding microenvironment comprises many different cell types that play specific roles in this complex functional network. The microenvironment accounts for nearly 80% of the breast volume and comprises an extracellular matrix and stromal cells including inflammatory cells, adipocytes, endothelial cells, and fibroblasts [120]. In the non-pregnant state, the mammary gland looks like a network of epithelial ducts that empty into the main lactiferous ducts. Epithelial cells’ secretory granules exocytosis releases several antimicrobial peptides inhibiting bacterial growth in the duct system [121]. With pregnancy, and because it must prepare for lactation, the epithelium markedly proliferates and differentiates. It expands to fill the gland, replacing the fat pad with milk-producing lobuloalveolar structures. When milk secretion stops, the mammary gland undergoes apoptosis of the lobuloalveolar cells generated during pregnancy and returns to its original ductal state. Milk SEVs containing MFG-E8 (milk fat globule-EGF (epidermal growth factor)-8) play an essential role in the recognition and engulfment of apoptotic epithelial cells by the neighboring phagocytic cells in the involuting gland [122]. As the epithelial cells are lost, the gland repopulates with adipocytes.
Stromal ECM, which mainly contains type I collagen, fibronectin, laminins, and glycoproteins, is a structural scaffold that maintains breast tissue integrity [123]. Fibroblasts regulate ECM deposition and differentiation of the neighboring epithelium [117,124]. Cell-matrix and cell-cell interactions play critical roles in developing the normal mammary gland, where SEVs can participate [122,125,126]. In the normal gland, SEVs regulate epithelial cell polarity as mammary epithelial cells are highly polarized [127,128]. Several studies have shown that epithelial SEVs that shed apically or basolaterally differ in cargo composition or concentration [129,130]. Distinct loading mechanisms for apical versus basolateral cargo have been suggested [131]. Therefore, polarized secretion of SEVs allows targeted delivery of specific SEV populations to stromal recipient cells due to the organized tissue architecture [128].

3.2. Exosome Role in the Maintenance of the Mammary Stem Cell

During female mammals’ sexually active life, the mammary gland continuously undergoes tissue remodeling [132]. During each cyclic pattern of ovarian activity, breast cells proliferate and form alveolar buds at the tertiary side branches, then regress in an ordered fashion [133]. Under pregnancy stimuli, lobuloalveolar differentiation takes place with breast epithelial expansion, which generates complex milk-secreting alveolar units, whose cells undergo terminal differentiation into specialized secretory cells in late pregnancy. Such a dynamic structure with high regenerative capabilities has suggested the existence within the breast of the renewable stem cell population [134]. Mammary stem cells (MaSCs) have been isolated and shown to be able to individually regenerate an entire mammary gland within six weeks in vivo while simultaneously executing up to ten symmetrical self-renewal divisions [135,136]. Localization of dormant MaSCs to the fat pad’s proximal region may indicate a specific microenvironment that resembles the MaSC niche [137]. Stromal fibroblasts appear to be a significant determinant of development in the mammary gland. Several fibroblast-derived factors have been implicated in transmitting signals to the epithelium, including morphogen ligands such as hedgehog or Wnt molecules [138,139]. Mutations within these key-signaling pathways can deregulate MaSCs from controlling regulatory signals, allowing them to develop precursor lesions [140]. Stromal SEVs can also participate in that regulation. Generally, mammary luminal cells do not have stem cell properties and cannot generate mammary glands when implanted into fat pads. SEVs derived from stem-like mammary basal cells can transfer mammary gland-forming abilities to luminal mammary epithelial cells [126]. Such SEVs release is regulated mainly by the presence of SEVs in the extracellular environment [105]. It has been shown to impose quiescence on residual hematopoietic stem cells in the leukemic niche [141].

3.3. Milk Is an Essential Source of Extracellular Vesicles

Breast milk, the most important nutritional source for infants, has many beneficial effects. It is rich in various nutrients and ingredients, including proteins, fats, carbohydrates, minerals, and vitamins, which can provide the energy necessary for growth and development in infancy [142]. It also contains many extracellular vesicles whose bilayered structure allows them to remain stable in the baby’s stomach until further absorption by intestinal cells [143]. Once absorbed, maternal EVs can enter the bloodstream and then infant tissues [144], where they will play different vital roles (for review [145]). They will have positive effects on the developing immune system [146,147] and play a role in metabolic regulation [148] and neural development of the newborn [149].
It is well known that pregnancy increases BC risk for all women for at least five years after parturition [150]. In such a context, breast milk extracellular vesicles may be necessary for BC as they may influence BC risk [151]. By promoting epithelial-mesenchymal transition (EMT), milk extracellular vesicles can increase the aggressiveness of both benign and malignant breast epithelial cells when the breast is remodeling, and the surrounding microenvironment is likely to be tumor-promotional [152]. Milk from healthy lactating women contains high levels of TGFβ2 in SEVs that have been evidenced to promote EMT, modifying both MCF7 breast cancer and MC10A breast benign cell lines morphology by disrupting cell-cell junctions and increasing filopodia formation [151].

4. Extracellular Vesicles Deregulation in Breast Cancer

4.1. Extracellular Vesicle and Cancer Stem Cells

Tumor initiation, therapeutic resistance, recurrence, as well as metastasis have also been associated with the concept of stemness and plasticity in BC [6,153,154]. A relatively rare self-renewal sub-population may drive epithelial cancers, multipotent cells, cancer stem cells (CSCs), or tumor-initiating cells (TICs). Unlike normal adult stem cells that remain constant in number, such cells can increase as tumors grow and give rise to progeny that can be either locally invasive or colonize distant metastatic sites.
As for any other adult stem cells, the properties of mammary stem cells (MaSCs) make them probable candidates for breast cancer initiation [155]. Self-renewal and asymmetric division are stem cells’ cardinal properties that are tightly regulated within the MaSC niche [156] and confer to MaSCs both preserved replicative capacity and resistance to differentiation. Indeed, MaSCs stemness acquisition occurs through the initiation of an epithelial-mesenchymal transition program that activates primary ciliogenesis, which then enables Hh signaling [157]. Mutations in such complex regulatory systems may induce the development of mammary TICS (MaTICs) [158,159] and the MaSCs neoplastic counterpart [160]. MaTICs that have already undergone epithelial–mesenchymal transition possess motility characteristics and can spread in foreign tissues to form a metastatic mass. Thus, MaSCs can harbor mutations over a prolonged life span [140], allowing them to be the true site of breast cancer initiation [161].
MaTICs represent a potential source of tumor heterogeneity. Their high plasticity associated with the random nature of mutations confers variable properties contributing to the considerable cellular heterogeneity observed in human breast cancers [162]. Either MaSCs or MaTICs share the capability to cross-communicate with their environment to maintain homeostasis. It allows the generation of mature breast functional cells throughout life without depleting the pool of stem cells [135,163,164]. The overabundance of microenvironmental stimuli received by the stem cell niche can support the observed phenotype MaTIC variability [165]. Aside from the numerous factors that can modulate the persistence of quiescent/slow-cycling cells in the niche, SEVs transfer figures [166]. Every tiny variation or modulation in SEVs delivery during the continuous crosstalk between CSCs and their surrounding microenvironment is critical and could induce significant deregulation and further tumor progression [167]. For example, PGE2/EP4 signaling controls the homeostasis of MaSCs through SEVs release regulation. MaSCs reprogramming can result from EP4-mediated stem cell property SEVs transfer between mammary basal and luminal epithelial cells [168]. MiR-130a-3p inhibits migration and invasion of MaSCs by regulating Rab5B [169]. Chemotherapy-induced BC cells secrete multiple SEVs miRNAs, including miR-9-5p, -195-5p, and -203a-3p, simultaneously targeting the transcription factor One Cut Homeobox 2 (ONECUT2), which induce CSC traits to BC cells, which has been associated with tumor refractoriness and progression [170]. In addition, BC cells prime mesenchymal stromal cells to release SEVs containing miRNAs such as miR-222/223, promoting dormancy in a subset of BC cells and conferring drug resistance [171]. Understanding the importance of SEVs transfer in that context is a crucial feature for future BC therapy [172].

4.2. Bidirectional Contributions of Extracellular Vesicles from Breast Tumor and Microenvironmental Cells to Breast Cancer Changes

The tumor microenvironment (TME) is a complex and dynamic network including cancer and stromal cells. Stress conditions such as hypoxia, starvation, and acidosis increase tumor cells’ SEVs release, leading to TME changes and expansion. Such specific behavior is the consequence of a complex combinatory of bioactive molecules present in SEVs [173]. Not only proteins (cytokines, etc.) or lipids but also different RNA forms could account for these critical changes. In breast tissue, miRNAs regulate the expression of cytokines and growth factors that can affect extracellular matrix composition and pave the way for pathogenesis [174].

4.2.1. Breast Cancer Cells-Derived Exosomes Transfer to Local Microenvironment

SEVs derived from BC cells can transform non-tumor breast ones. SEVs derived from MDA-MB-231 BC cells induced epithelial-mesenchymal transition (EMT) [175] while those derived from triple-negative breast cancers (TNBC) HCC1806 cell line induced proliferation and drug resistance in MCF-10 breast epithelial cells [176].
SEVs could transfer miR-370-3p from BC cells to normal fibroblasts, facilitating their activation through CYLD down-regulation and further NF-ζB signaling pathway [177], leading to cancer-associated fibroblasts (CAFs) with protumorigenic and proangiogenic properties [178] (Figure 3). miR-9 was also found to convert normal fibroblasts into CAFs, and its overexpression also identified a signature of different genes related to cell motility and extracellular matrix organization [179]. BC cell SEVs encapsulated miR-105 can mediate metabolic reprogramming of CAFs through Myc signaling [180].
TNBCs are highly infiltrated by tumor-associated macrophages (TAMs). TNBCs release SEVs and soluble molecules that promote, via TLR2 and TLR3 Toll-like receptors, monocyte differentiation toward TAM fates to phenocopy the tumor and rewire the microenvironment [181]. SEVs, combining either surface CSF-1 promoting survival or cargoes promoting cGAS/STING pathway, specifically promoted macrophage differentiation into proinflammatory TAMs bearing an interferon response signature [182]. Delivery of BC cell-derived SEVs containing miR-138-5p downregulates KDM6B expression inhibiting M1 and stimulating M2 polarization [183]. Likewise, lncRNA BCRT1 secretion mediated by BC exosomes promoted M2 polarization, further accelerating BC progression [184]. This SEVs-induced pro-survival macrophage differentiation is driven through IL-6 receptor beta/glycoprotein 130/STAT3 signaling pathway [185].
Additionally, surrounding endothelial cells (EC) can be activated by BC cell-secreted SEVs. Exosomal Annexin II (AnxA2) transfer from BC cells has been shown to promote EC angiogenesis [186].
Lastly, BC cells can also interact with adipose tissue [187] through exosome transfer. Normal adipocytes are driven into cancer-associated adipocytes by tumor cells [188] through SEVs transfer of oncomiRs [189]. BC cells-derived EVs can also convert adipose tissue-derived MSCs to myofibroblasts [190].

4.2.2. The Microenvironment Produces Exosomes That Could Transform Breast Cancer Cells

In response to BC cells, TME modifications induce SEVs-driven stromal cell response, resulting in tumor changes by further modifying BC cells [191]. This continuous dual SEVs-driven interplay between stromal cells and BC cells is central in tumor behavior as it may drive either tumor cell proliferation or migration. Among TME, CAFs, ECs, and infiltrating TAMs are likely to be the major cell types interacting with BC cells or within the TME through SEVs signaling.
CAFs are well known to play a pivotal role in controlling cancer cell invasion and metastasis, immune evasion, angiogenesis, and chemotherapy resistance [192]. CAF-derived exosomes carrying miR-181d-5p can promote proliferation, invasion, migration, and EMT and inhibit BC cell’s apoptosis by downregulating CDX2 and its downstream target HOXA5 [193]. BC cells’ endocytosis of CAFs SEVs miRs -21, -378e, and -143 increased their capacity to form spheres, stem cell and EMT markers expression, and anchorage-independent cell growth [194]. Transfer of CAFs p85α-deficient SEVs carrying the Wnt10b protein into BC cells induced EMT [195]. CAFs SEVs could also reprogram BC cell metabolism by modulating pyruvate kinase PKM2 expression through the enrichment of exosomal noncoding RNA [196]. Once metabolically reprogrammed, miR-105 transformed CAFs promote glutamine and glucose metabolism to feed adjacent BC cells [180]. Lastly, SEVs transfer results in CAFs activation through miR-146a/TXNIP axis to activate the Wnt pathway, which in turn enhances the invasion and metastasis of BC cells [197].
SEVs’ transfer from EC drives a cadherin switch in BC cells that favors further intimate contacts between EC and BC cells [198]. Blocking IL-3R-alpha suppresses EC SEV-induced angiogenesis stimulation by targeting the Wnt/β-Catenin pathway [199].
Once transformed, TAMs can also transfer SEVs to BC cells [200]. A recent report shows that TAMs SEVs-driven noncoding RNA molecules transfer will boost BC cell proliferation and direct their phenotype and metabolic changes to progression and metastasis [201].
Adipocytes are, by mass, the preponderant non-malignant cell type in BC TME. Adipocyte tissue (AT)-derived SEVs can also enhance growth, motility, and invasion, induce stem cell-like properties, and specific EMT features in estrogen receptor (ER)-positive and TNBC cells [202]. Among the AT SEVs activated signaling pathways in BC cells are Hippo [203], HIF-1α [202], ERK [204], Wnt/β-catenin [205], JAK/STAT3 [206], PI3K/AKT, and TGFbeta/Smad [207] (For review, [208]).

4.2.3. Local Inflammation at the Tumor Site and Extracellular Vesicles

Chronic inflammation is likely to be an essential driver in triggering tumor progression and metastasis [209]. In such process installation, SEVs are likely to play an important role [210]. Triple-negative TNBC cells release SEVs and soluble molecules promoting specific monocyte differentiation toward proinflammatory macrophages bearing an interferon response signature [182]. BC tumor-derived SEVs can induce an M1 proinflammatory response in macrophages through the activation of NFκB, which stimulates the production of inflammatory cytokines including GCSF, IL-6, IL-8, IL-1β, CCL2, and TNF-α [211]. NF-κB is a significant regulator of inflammation, and constitutive activation of NFκB is often observed in BC cells and associated with an aggressive phenotype. This M1 activation of NFκB, but also p38 MAPK and STAT3 pathways, seems to be triggered by high-level annexin A2 containing SEVs [186].

4.3. Promotion of Tumor Expansion

Accumulated genetic and epigenetic changes often activate the expression of oncogenes while silencing tumor suppressors during carcinogenesis [212]. BC genomic instability leads to several protooncogene mutations affecting multiple signaling pathways [213]. Cell cycle pathways (gain of function mutations of cyclin E, cyclin D, and CDK2/4/6) are transformed in about 50% of all BC types [214]. MAPK signaling is greatly amplified in 80% ERBB2-positive BC [215]. PI3K pathway is altered in more than 60% of luminal A BC [214] while 90% TBNCs undergo p53 inactivation [216]. SEVs released by CXCR4-positive BC cells increase the oncogenic potential of tumor cells in mice [217]. Hypoxic BC cells produce a high amount of SEVs containing long non-coding lncRNAs SNHG1, which, when upregulated, acts as an oncogene [218]. Their transfer, targeting the miR-216b-5p/JAK2 axis, promotes growth in vivo by upregulation of the JAK2/STAT3 pathway [219]. BC gain of function p53-containing small SEVs convert surrounding tumor microenvironment fibroblasts to cancer-associated ones [220]. Interestingly, many oncogenes, especially MYC and AURKB, can regulate either SEVs’ biogenesis or release in BC cells [221].

4.4. Cancer Metabolism Reprogramming

Throughout the natural history of cancer, tumor cells should unveil high metabolic plasticity to adapt to continual changes within the tumor and surrounding environment [222]. Tumor cell proliferation must continuously adjust their metabolism to meet the highest nutrient capacity to fulfill enhanced biosynthetic and bioenergetics demands. In normal cells, glycolysis and mitochondrial oxidative phosphorylation (OXPHOS) cooperate to produce energy. In BC, as mitochondrial ATP production is generally impaired, tumor cells enhance glycolytic glucose consumption to get sufficient ATP, thus generating a high lactate content even in aerobic conditions (“Warburg effect”) [223,224]. While increasing evidence associates metabolic reprogramming to OXPHOS and subsequently enhanced glutaminolysis with the induction and maintenance of the epithelial–mesenchymal transition program [225], the use of mitochondrial metabolism in BC cells migration and invasion is still controversial [226]. Mammary tumor-initiating cells (MaTICs) seem more dependent on OXPHOS, producing less lactate [227]. Triple-negative TNBCs mostly rely on glycolysis [228], while reprogramming to OXPHOS is associated with a higher risk of recurrence and death [229]. High lactate production and secretion induce tumor microenvironment acidification promoting immune surveillance escape and metastasis [230].
Recently, it has been evidenced that BC tumor microenvironment metabolism can largely modulate cancer cell progression [231]. Cancer-associated fibroblasts (CAFs) can provide metabolites that will facilitate tumor cells’ ATP production. Lactate, exported through CAFs MCT4 lactate shuttle then uptaken through cancer cells MCT1 lactate transporter, could be used to fuel surrounding cancer cells, a process called “reverse Warburg effect” [232,233]. Adipocytes’ free fatty acid (FA) secretion followed by free FA CD36 uptake promotes BC cells progression [234]. Either FA synthesis and FA oxidation or glutamine and serine metabolisms all increase in tumor cells as lipids, amino acids, and nucleotides are strongly required for their multiplication [235,236]. BC cells and surrounding tumor microenvironment cells can shed SEVs that will modulate cancer cell metabolism and play a role in their proliferation. SEVs can contain metabolites and metabolism enzymes that can modulate cancer cells’ metabolism. For example, GLUT-1 glucose transporter was enriched in BC cells SEVs [237]. MDA-MB-231 BC cell line SEVs increase the peripheral blood mononuclear cells’ expression of GLUT1 and hexokinase HK2 genes, which are effective in the glycolysis pathway [238]. BC cell-secreted miR-122 reprograms glucose metabolism in the premetastatic niche to promote metastasis [239]. MDA-MB-231 BC cell-derived SEVs led to pyruvate kinase M2 (PKM2) phosphorylation in MCF7 cells that acquired a more aggressive phenotype, which resulted in increased aerobic glycolysis and cell proliferation [240]. Exosomal miR-105 from BC cells can alter the glucose metabolism of stromal cells and thus promote the growth of cancer cells under nutrient-deprived conditions [241]. In addition, miR-105 combined with miR-204 targets RAGC to regulate mTORC1 upon amino acid stimulation. Affected fibroblasts exhibit reduced mRNA translation and selective protein synthesis [242]. miR-144 containing SEVs downregulates the MAP3K8/ERK1/2/PPARγ axis, thus inducing beige/brown differentiation, while miR-126 remodels metabolism by disrupting IRS/Glut-4 signaling and activating the AMPK/autophagy pathway in resident adipocytes [189]. Exosomal miR-155 promotes lipolysis in adipocytes and facilitates an aggressive phenotype of BC-derived tumor cells [243].
In parallel, SEVs from tumor microenvironment cells can modulate BC cells’ metabolism. SNHG3 knockdown in CAF-secreted exosomes suppressed glycolysis metabolism and cell proliferation by the increase of miR-330-5p and decrease of pyruvate kinase PKM expression in tumor cells. SNHG3 functions as a miR-330-5p sponge to positively regulate PKM expression, inhibit OXPHOS, increase GLYC, and enhance BC cells’ proliferation [196]. Metabolic remodeling in cancer-associated adipocytes surrounding BC cells enhances tumor aggressiveness by promoting cancer cell survival and proliferation through SEV production [244]. As for CAFs, miR-105 can also activate MYC signaling in cancer-associated adipocytes to induce a metabolic program secreting energy-rich metabolites to fuel neighboring cancer cells [241].

4.5. Angiogenesis Induction

Angiogenesis is an essential feature for tumor proliferation and further metastasis. The uptake of tumor-derived SEVs by ordinary endothelial cells activates angiogenic signaling pathways in endothelial cells (ECs) and stimulates new vessel formation [245]. Once internalized, SEVs are immediately directed to the perinuclear zone and actin filaments-rich area. When tubules are formed, SEVs move to the cell periphery and enter advanced pseudopods [246]. After complete remodeling, adjacent ECs probably transport SEVs to neighboring ECs and other cells within the tumor microenvironment [111]. In hypoxic conditions, BC cells can secrete angiogenic factors, such as VEGF-A, inducing ECs migration and tumor angiogenesis [247]. Aside from Notch signaling [248] and angiopoietins [249], VEGF is one of the more potent angiogenesis promoters, thus behaving as an important mitogen with high specificity for ECs [250]. Studies have reported that SEVs released from hypoxic tumors are more likely to cause angiogenesis and vascular leakage; hypoxia gradually promoting, through HIF-1α signaling, BC cells SEVs release [251]. SEVs transfer with miR-210 from hypoxic BC cells to cells into the tumor microenvironment-induced expression of vascular remodeling-related genes, such as Ephrin A3 and PTP1B, to promote angiogenesis [252]. The same miR-210 and a set of other angiogenic miRNAs are enriched in SEVs released by metastatic BC cells, a secretory process regulated by neutral sphingomyelinase 2 (nSMase2, SMPD2). These SEVs, once transferred to ECs, enhance the capillary formation and migration capability [253]. Docosahexaenoic acid (DHA) has potent anticancer properties, mainly through VEGF suppression [254]. A recent report showed that DHA increased the expression of anti-angiogenic miRNAs (i.e., miR-34a, miR-125b, miR221, and miR-222) while decreased levels of proangiogenic miRNAs (i.e., miR-9, miR-17-5p, miR-19a, miR-126, miR-130a, miR-132, miR-296, and miR-378) in SEVs derived from DHA-treated BC cells [255]. SNHG1 enclosed in BC cells SEVs induces angiogenesis via regulating the miR-216b-5p/JAK2 axis [219]. circHIPK3 enhanced MTDH expression in the EC by sponging miR-124-3p, favoring endothelial tube formation [256]. Another miR, miR-22-3p, mediated tumor vessel abnormalization by suppressing transgelin, thus promoting tumor budding and BC progression in vivo [257].
Aside from the various forms of SEV RNAs involved in its promotion, SEV-specific proteins can also stimulate angiogenesis. STIM1 promotes angiogenesis by reducing exosomal miR-145 in BC MDA-MB-231 cells [258]. Annexin II (AnxA2), a Ca2+-dependent phospholipid-binding protein associated with the plasma membrane, is one of the most expressed proteins in SEVs [259]. BC-derived SEVs transfer proangiogenic AnxA2 to ECs and induce angiogenesis by the tPA-dependent increase in plasmin generation [186]. Ephrin-A2 (EPHA2) was also rich in highly metastatic BC-derived exosomes and confers a proangiogenic effect [260]. Heparanase helps drive SEVs secretion and alters exosome composition (increase in matrix metalloproteinase-9 (MMP-9), VEGF, hepatic growth factor 2 (HGF2), and receptor activator of nuclear factor κ-B ligand (RANKL)) that impact both tumor and host cell behavior [261]. Lastly, EC-derived SEVs themselves can play a role in BC cells. ECs SEVs contained soluble and membrane-anchored forms of VE-cadherin that drive a cadherin switch in BC cells and neo-expression of VE-cadherin [198]. On the other hand, it has been previously reported that mesenchymal stromal cells MSC-derived SEVs negatively modulate angiogenesis by down-regulating BC cells’ VEGF synthesis through miR-16 transfer [262]. Additionally, MSC-derived SEVs enrichment with miR-100 suppresses angiogenesis in vitro by VEGF down-regulation through mTOR/HIF-1α)/VEGF signaling axis modulation [263]). Such findings emphasize SEVs’ multifaceted role in tumor-to-stroma communication within the TME.

4.6. Immune Evasion

Before the clinical presentation, most malignant cells are eliminated by immune surveillance through combined stimulation of innate and adaptive immune responses [264]. Nevertheless, BC cells, like other cancer cells, must evade immune control, a prerequisite in the transition from preinvasive to potentially lethally invasive disease [265]. In the early steps of tumor development, host immune factors play a crucial role in rejecting cancer cells [266]. Thus, clonal evolution patterns during progression will depend on the immune context [267]. Some progressing clones become immune privileged, despite tumor-infiltrating lymphocytes, while immunoedited tumor clones are eliminated [268]. To evade the immune system, tumors release immunosuppressive cytokines (e.g., TGF-β, interleukins IL8, IL6, IL10, etc.) and skew the tumor microenvironment to a more immunosuppressive one through either inhibiting CD8+ T cells, NK cells, dendritic cell maturation or increasing Tregs and tumor-associated macrophages (TAMs) [269]. SEV signaling in BC has been shown to play a crucial role in the crosstalk between immune and cancer cells [270]. Tumor-derived SEVs interacting with immune cells deliver negative signals to these cells and interfere with their antitumor functions [271,272].
The immunosuppressive nature of BC SEVs was confirmed in vitro where it promoted T-cell exhaustion and NK-cell cytotoxicity [273] (Figure 3). Hypoxia enhances SEVs secretion by BC cells, which acts to suppress T cell proliferation via TGF-β [274]. In the meantime, tumor-derived SEVs also promote Treg expansion and increase their immunosuppressive functions [275]. SEVs from 4T1 murine BC cells blocked the differentiation of myeloid precursor cells into CD11c+ DCs and induced cell apoptosis. They also drastically decreased CD4+IFN-γ+ Th1 differentiation but increased the rates of Treg cells [276]. While promoting the in vitro expansion of CD4(+)CD25(+)FOXP3(+) Treg cells and enhancing their suppressor activity, BC cells Mage 3/6 positive SEVs also inhibited signaling and proliferation of activated CD8(+) but not CD4(+) T cells and induced apoptosis of CD8(+) T cells [275]. BC cells SEVs SNHG16 lncRNA induced CD73 + γδT1 cells to act as immunosuppressive regulatory T cells by activating the TGF-β1/SMAD5 pathway [277].
Human MDA-MB-231–derived SEVs induce M2-type macrophage polarization (upregulation of CD206 and arginase-1), supporting enhanced tumor growth and axillary lymph node metastasis in an orthotopic triple-negative TNBC model [278]. miR-34a in triple-negative TNBCs mediate M1 polarization, while antagomiR-34a promotes M2 plasticity [279]. miR-503 in BC patients is vital in promoting brain metastasis by programming the microglia through M1 to M2 macrophage polarization induction [280]. Additionally, BC cells’ SEVs LncRNAs repertoire correlates with macrophage polarization [281]. Specifically, SEVs lncRNA BCRT1 promoted M2 polarization of macrophages, further accelerating BC progression [184]. The combination in TNBC SEVs of surface CSF-1 promoting survival and cargoes promoting cGAS/STING or other activation pathways led to the differentiation of this particular macrophage subset [182].
Cancer-associated fibroblasts (CAFs)-derived exosomes suppress immune cell function in BC by regulating PD-L1 levels in BC cells via the miR-92/LATS2/YAP1 pathway [282]. Lin28B promotes lung metastasis of BC cells by building an immune-suppressive pre-metastatic niche. Lin28B enables neutrophil recruitment and N2 conversion. The N2 neutrophils are then essential for immune suppression in the pre-metastatic lung by PD-L2 up-regulation and a dysregulated cytokine milieu [283]. Lastly, tumor-derived SEVs inhibited NK cell immunity using murine mammary (TS/A) tumor cell lines. TS/A SEVs are taken up by NK cells and account for decreased cytotoxic activity. Not only TS/A SEVs but also SEVs from human MDA231 or murine 4T1 BC cells could significantly block the proliferation of NK cells induced by IL-2 [284].

4.7. Metastatic Spread Induction and Secondary Settlement

Metastasis, which causes over 90% of BC-related deaths, behaves as a cascade comprising local invasion, intravasation, survival in the circulation, premetastatic niche modeling, and extravasation, and then metastatic niche colonization [285].

4.7.1. Extracellular Vesicles and Epithelial to Mesenchymal Transition of BC Cells

Within the primary tumor, epithelial-to-mesenchymal transition (EMT) that confers enhanced mobile capabilities to tumor cells is likely to be one of the primary metastatic events. In BC, EMT activation has been shown to increase stemness [286], with most of the hematogenous circulating cancer cells harboring a mesenchymal phenotype [287]. Tumor-derived SEVs can facilitate EMT [288,289,290]. SEVs derived from mesenchymal stromal cells contained several molecules able to induce EMT such as well-known inducing proteins such as transforming growth factor-beta (TGF-β), hypoxia-inducible factor-alpha (HIF1α), or β-catenin as well as miRNAs, lncRNAs, and circRNAs [291]. CAF-derived exosomal miR-181d-5p can regulate CDX2 and HOXA5 in BC cells, thereby promoting their EMT [193]. Increased concentrations of MiR-9, miR-424, and miR-155 in SEVs led to BC cell EMT and aggressiveness [243,292,293]. SEVs are associated with pro-metastatic phenotype reprogramming in recipient surrounding cancer cells [294]. SEVs from human adipose-derived mesenchymal SCs promote migration through the Wnt signaling pathway in a BC cell model [205].

4.7.2. Extracellular Vesicles Impact on Extracellular Matrix Disruption

The extracellular matrix (ECM) is also an essential regulator of BC progression [295]. BC SEVs can increase cancer cell invasion by containing mediators of cancer progression and critical factors in tissue remodeling, a prerequisite for seeding [296,297]. In that process, ECM stiffening due to excess deposition and crosslinking of collagen dramatically influences tumor behavior and fate by orienting fibers, thus facilitating metastatic cell intravasation [298]. Stiff ECM promotes SEV secretion in a YAP/TAZ pathway-dependent manner and triggers BC invasiveness using thrombospondin-1 (THBS1) as a master player [299]. BC cell-derived SEVs can also cargo ECM degradation enzymes such as MMPs, etc., as well as their regulators [290,300]. Tumor-derived SEVs transferred surface-bound proteases such as glycosidases to cleave ECM components, resulting in ECM remodeling and facilitating tumor development [296,301]. Interestingly, silencing Rab GTPases that tune biogenesis and secretion of pro-metastatic SEVs in BC cells, upregulate the levels of MCAM and CD146 adhesion molecules and limit BC metastasis [302].
To initiate the metastatic process, BC cells will also recruit and educate stromal cells to induce cancer-associated fibroblasts (CAFs), tumor-associated macrophages (TAMs) with the immune-suppressive M2 phenotype, and endothelial cells that promote tumor angiogenesis [303]. SEVs are likely to be critical players in this mandatory recruitment that contributes to the ability of BC cells to metastasize [304,305]. CAF-derived SEVs can promote BC cell motility through two independent mechanisms involving Wnt [306] or Notch [307] signaling in the cancer cells. Tumor SEVs mediate the migration of MDSCs and contribute to the metastasis of murine BC cells (4T1 cells) to the lung in a CCL2-dependent manner [308].

4.7.3. Extracellular Vesicles and BC Cells Spread

Once the extracellular matrix is disrupted, the distant spread can then arise in two steps. The first concerns local tumor cell dissemination, where epithelial cells migrate through the tumor microenvironment at the front of the tumor through the generation of membrane protrusions (invadopodia) and basal lamina break-in [309]. It was clearly shown for BC using an injection of MTLn3 cells. This highly invasive rat mammary adenocarcinoma cell line forms invadopodia in vitro into the mammary gland of immunocompromised mice and rats and allows them to form tumors [310]. Cancer-associated fibroblasts CAFs-SEVs enhance this BC cell protrusive activity and motility via Wnt-planar cell polarity signaling [306]. The second involves vascular disruption to allow tumor cells hematogenous spread. Tumor SEVs can increase vascular permeability to promote the extravasation of circulating tumor cells (CTCs). Both exosomal miR105 and miR-939 secreted by metastatic BC cells are involved in VE barrier destruction, thereby increasing vascular permeability and promoting distant metastasis [180,311]. Identically, metastatic BC cells facilitate brain metastasis by releasing miR-181c-containing SEVs capable of destroying the blood-brain barrier [312].

4.7.4. Extracellular Vesicles, Pre-Metastatic Niche, and Secondary Organ Settlement

As SEVs are not limited to the local tumor microenvironment, they can also cargo “tumor-nourishing” environments at distant sites to encourage metastatic settlement. SEVs educate a metastatic microenvironment, commonly defined as the pre-metastatic niche allowing circulating tumor cells (CTCs) to find a suitable environment in which they can settle and then proliferate. Such niche generation is characterized by local tissue inflammation, immune suppression, stromal cell activation, and ECM remodeling. Pre-metastatic niches are characterized by key tissue architecture, composition, and metabolism modifications, facilitating CTCs’ arrival, survival, and further expansion [313]. SEV-mediated intercellular interactions can generate a pro-metastatic tumor microenvironment [314]. By modifying glucose utilization by recipient pre-metastatic niche cells, BC-derived extracellular miR-122 can reprogram systemic energy metabolism to facilitate metastatic progression [239]. MiR-940 overexpression induced in MDA-MB-231 BC cells has been shown to induce extensive osteoblastic lesions in mice by facilitating the osteogenic differentiation of host mesenchymal cells [315].
An important characteristic of tumor cells relies on their capacity to colonize preferentially specific organs (organotropic metastasis) that are often determined by anatomic aspects. According to the Paget “seed and soil” theory [316], SEVs can even be considered as the “soil conditioner in BC metastasis” [317,318], leading to an inflammatory and mechanical niche promoting survival and colonization of immigrant CTCs. Indeed, bone marrow lesions were observed in mice bearing mammary cancer far before the arrival of tumor cells [319]. Integrin (ITG) SEVs repertoire seems to drive organ-specific metastasis, ITG α6β4 and αvβ3 on the surface of BV SEVs increasing lung metastasis [320]. During the establishment of an inflammatory environment in organs to which tumors will metastasize, SEVs contribute to the upregulation of proinflammatory cytokines and inflammation-activating factors, as well as the recruitment of immune cells to the pre-metastatic niche. BC-derived SEVs containing CCL3, CCL27, and other molecules are found to remodel the bone microenvironment, characterized by stimulating osteoclastogenesis and angiogenesis [321]. BC and lung tumor-derived SEVs containing Cell Migration-Inducing and hyaluronan-binding Protein (CEMIP) could induce a proinflammatory vascular niche by upregulating cytokines Ptgs2, TNF, and CCL/CXCL cytokines to promote brain metastasis [322]. Annexin A2 released by BC cells’ SEVs can induce macrophage-mediated activation of either p38 MAPK, nuclear factor κB (NF-κB), or STAT3 pathways, thus increasing IL-6 and tumor necrosis factor (TNF)-α secretion, thereby contributing to the formation of a premetastatic inflammatory microenvironment in distant organs such as the lung and brain [186]. Interestingly, once in the brain, BC cells’ survival is increased by SEV-encapsulated miR-19a released by astrocytes that act by decreasing PTEN expression [323].
Cancer cell SEVs can reprogram resident cells like in primary tumors to promote metastatic niche achievement and attract newly released CTCs. BC-derived exosomal microRNA-200b-3p uptaken by alveolar epithelial type II cells (AEC II) induces the high expression of CCL2, S100A8/9, MMP9, and CSF-1 in the lung to recruit myeloid-derived suppressor cells (MDSCs) and promote inflammatory pre-metastatic niche formation [324]. SEVs secreted by highly metastatic murine BC cells inhibit antitumor immune responses in premetastatic organs, directly suppressing T-cell proliferation and NK cell cytotoxicity [325]. BC cells SEVs remodeled lung parenchyma via a macrophage-dependent pathway to create an altered inflammatory and mechanical response to tumor cell invasion [326]. Immune cells can also play an important role in these distant organs [327]. BC-derived SEVs containing ANXA6 are targeted to the lung and activate the CCL2-CCR signaling axis, thus recruiting monocytes, which then differentiate into macrophages at this future site of metastasis [328]. These metastasis-associated macrophages (MAMs) have been first described in mouse models of BC lung metastasis [329]. At both bone and lung sites, MAMs promote BC cell extravasation, seeding, and metastatic outgrowth [330,331].

4.8. Cancer Cells Dormancy

Metastatic disease can occur years or even decades after the first diagnosis and subsequent treatment, suggesting that cells initiating recurrence are often long-lived and able to reactivate proliferation after long latency periods (also referred to as clinical dormancy) [332]. In BC, late recurrences (>5 years) account for most of the deaths among patients [333]. It is likely that these specific metastatic tumor cells exit the cell cycle and remain in a growth-arrested state [334]. Dormant tumor cells are commonly referred to as slow-cycling cancer stem cells that combine quiescent properties with tumor-initiating and chemoresistant properties, which favor later relapse and for the formation of metastases [335,336]. Dormant BC cells exhibit a distinct gene expression signature from metastatic ones regardless of the metastatic site [337]. When circulating tumor cells (CTCs) first extravasate from the vessels, they may reside in a niche surrounding the microvasculature, the perivascular niche (PVN) [338] that comprises resident hematopoietic cells, endothelial cells, and mesenchymal stromal cells (MSCs). Evidence has accumulated over recent years that the PVN in BC orchestrates CTC dormancy, principally responsible for cell survival and growth arrest [339]. CTCs may receive intrinsic factors (microenvironmental factors and signaling molecules) relevant to dormancy [340]. In bone marrow representing a niche for BC cell homing, SEVs from surrounding MSCs contain specific miRNAs that drive metastatic BC cells to dormancy. Gap junction-mediated import of microRNAs, including miR-222/223, mir-127, and mir-197 from bone marrow stromal cells, have been shown to elicit cell cycle quiescence in BC cells [341]. Interestingly, stromal SEVs are likely to also play an essential role as those containing miRNAs, such as miR-222/223, promote quiescence in a subset of BC cells [171]. Additionally, they also contribute to the dormancy of BC cells by reducing either CXCL12 levels [342] or targeting ERK1/2 signaling via miR-148a-3p [343]. Overexpression of MSCs-SEVs miR-23b in highly metastatic BC BM2 cells induced dormant phenotypes through the suppression of a target gene, MARCKS, which encodes a protein that promotes cell cycling and motility [344]. SEVs-enclosed miR-205 and miR-31, targeting the ubiquitin-conjugating enzyme E2N (UBE2N/Ubc13) and downregulating its activity, induced dormancy in MDA-MB-231 cells [345]. Interestingly, BC cells primed with MSCs SEVs were more highly resistant to chemotherapy [346]. In addition, SEVs from differentially activated macrophages influence the dormancy or resurgence of BC cells within the bone marrow stroma. M2 metastasis-associated macrophages (MAMs) form gap junctions with CSCs, resulting in cycling quiescence, reduced proliferation, and carboplatin resistance. Activation of M2 MAMs via the toll-like receptor 4 (TLR4) switched to the M1 phenotype can occur directly or indirectly through the activation of MSCs. Thus, M1 MAM SEVs activated NFκB reverse quiescent BC cells to cycling cells [347]).

4.9. Resistance to Therapy

BC is a heterogeneous disease in which each patient has individual characteristics that must drive treatment choices. In such a context, searching for new markers to improve the diagnosis and prognosis and achieve a better treatment response is mandatory. Currently, strategies for treating BC depend on the tumor subtype, and the selected treatments are directed to specific targets that are functionally altered in each cancer subtype. Conventional treatments for the management of BC patients have included endocrine therapy, targeted immunotherapy, and chemotherapy, all of these agents being used in adjuvant, neoadjuvant, and metastatic settings [5]. However, despite the improvement and diversification of therapeutics for BC patients and the emergence of new drugs during the last years, resistance to treatment remains a deadlock for women with an advanced BC for whom medicines no longer work.

4.9.1. Resistance to Hormone Therapy

Nearly 80% of BC are estrogen receptor positive (ER+) [348], the vast majority of them being initially dependent upon the activation of ER by estrogens [349]. Because of the importance of the estrogen-ER axis in breast tumorigenesis, the main treatment options for these patients are still endocrine therapies such as aromatase inhibitors, selective modulators of ER activity, or selective ER down-regulators. Among all patients with BC who have hormone receptor-positive tumors, 84% receive hormonal therapy [350]. Nevertheless, the major challenge in treating ER+ BC is to overcome endocrine resistance, whose mechanisms can be very complex [351,352,353,354]. SEV secretion can be involved in these processes [355]. SEVs from tamoxifen-resistant MCF7 (MCF7TR) BC cells transfer resistance inducing epithelial–mesenchymal transition and resistance to apoptosis to wild-type ones [356]. Several SEVs transfers of noncoding RNAs conferring tamoxifen resistance have been reported: miR-221/222 that confers stem cell-like properties [357], miR-9-5p [358], circ_UBE2D2 by binding to miR-200a-3p [359], miR-22 [360], and UCA1 LncRNAs [361]. SEV-mediated transfer of mitochondrial DNA (mtDNA) has been shown to promote an escape of BC cells from metabolic quiescence and led to hormone therapy resistance both in vitro and in vivo [362]. Aside from tamoxifen treatment, enhanced SEVs production has also been reported in aromatase inhibitor-resistant BC cells [363].

4.9.2. Resistance to Chemotherapy

Chemotherapy is generally used as neoadjuvant or adjuvant treatment in stage I-III BC (Miller 2022), while 60% of women diagnosed with metastatic disease (stage IV) most often receive radiation and chemotherapy alone [364]. Anthracyclines and taxane derivatives are two agents commonly used in the treatment of BC, but the emergence of chemoresistance often limits their efficacy. SEVs can be involved in the development of resistance. BC cells could export the chemotherapeutic drug doxorubicin (DOX) into the extracellular medium through vesicle formation, limiting its action [365]. SEVs isolated from the HCC1806 triple-negative TNBC cells can induce proliferation and drug resistance in the non-tumorigenic MCF10A breast cells [176].
Many studies have shown that resistance can arise through SEVs-mediated horizontal transfer of membrane-embedded drug efflux pumps to sensitive cancer cells. SEV-like structures containing the ATP-binding cassette (ABC) transporter protein ABCG2 have been reported to be increased in a variant of MCF-7 cell line with 20-fold resistance to mitoxantrone [366]. SEV delivery of P-gp (ABCB1) transporter was suggested to play an essential role in transferring drug resistance from DOX-resistant cells to drug-sensitive BC ones [367]. The transient receptor potential channel (TrpC) seems to play a crucial role in the upregulation of P-gp in drug-resistant BC cells [368]. Interestingly, SEVs from DOX-resistant MCF-7 cells were found to transfer TrpC5 (and also P-gp) to recipient human microvessel endothelial cells (HMECs) and further induce de novo expression of P-gp in these cells [369]. In sensitive MCF-7 cells, TrpC5-containing SEVs internalization led to Ca2+ influx through TrpC5s, which resulted in the upregulation of P-gp [370]. Interestingly, a strong correlation in nonresponsive tumors was observed between treatment resistance and increased TrpC5 expression by immunohistochemistry on BC patients’ tissues [371]. Ubiquitin carboxyl-terminal hydrolase-L1 UCHL1 (UCH-L1) was also found to upregulate P-gp expression by activating the MAPK/ERK pathway. UCH-L1-containing SEVs secreted by DOX-resistant human BC cells were taken up by DOX-sensitive human BC cells in a time-dependent manner and ultimately contributed to the chemoresistance phenotype [372]. SEVs transfer can confer neither DOX nor docetaxel (DTX) nor paclitaxel (PTX) resistance. PTX treatment induced the secretion of survivin-enriched SEVs from MDA-MB-231 cells, which highly promoted the survival of PTX-treated fibroblasts and SK-BR-3 cells [373].
Aside from protein transfer, noncoding RNAs have been involved in chemotherapy resistance induction. Resistant cells SEVs miR-100, miR-222, and miR-30a transfer confer resistance to wild-type ones [374]. Transfer of miR155 by inhibiting tetraspan 5 and promoting stemness [375,376], miR1246 by inhibiting Cyclin-G2 induced BC cells resistance [377], and miR-887-3p by targeting BTBD7 and activating the Notch1/Hes1 signaling pathway [378] can induce chemoresistance. Adaptation of cancer stem-like cell traits through SEVs transfer, including miR-9-5p, miR-195-5p, and miR-203a-3p targeting ONECUT2 has been shown to confer resistance [172]. SEVs derived from cisplatin-resistant MDA-MB-231 cells are characterized by a high expression of miR-423-5p that can be transferred to non-resistant cells [379]. LncRNA H19 was strikingly overexpressed in DOX-resistant BC cells and encapsulated into SEVs to transfer drug resistance. Similarly, the downregulation of H19 reversed DOX chemoresistance in sensitive BC cells [380].
Up to date, to overcome chemoresistance in BC treatment, several large-scale validation studies have been performed to determine the exosomal protein and miRNA expression profiles in drug-resistant BC after chemotherapy to find new potential markers and to better understand the transmission of SEVs-mediated chemoresistance [381,382,383,384,385].

4.9.3. Resistance to Radiotherapy

Resistance transfer through SEVs seems to also be a potent mechanism that confers resistance to radiotherapy [386]. SEVs derived from radioresistant cells can increase cell viability and colony formation in naïve recipient ones and increase their radiotherapy resistance [387]. Cargo from irradiated cell-derived SEVs was distinct from non-irradiated cells, indicating alterations in the exosomal formation system [388]. SEV levels of proteins such as PERP, GNAS2, GNA13, ITB1, and RAB10 correlate with BC cell trastuzumab response [389]. X-ray irradiation activates SEV biogenesis and secretion in a dose-dependent manner in MCF7 cells and induces their resistance to radiotherapy [390]. Such effects of radiation not only concern BC cells but also surrounding tumor microenvironment ones. In a mouse BC model, SEVs derived from irradiated cells elicited immune responses of tumor-specific CD8+ T cells and inhibition of tumor size [391]. Likewise, radioresistant SEVs stimulate tumor-supporting fibroblast activity, facilitating tumor survival and promoting cancer stem-like cell expansion [392].

4.9.4. Resistance to Targeted Therapy

As human epidermal growth factor receptor 2 (HER2) overexpression is often associated with BC poor prognosis, HER2-targeted therapy has been developed and achieved excellent efficacy in treating HER2+ BC [393]. Whereas trastuzumab generally has an excellent initial clinical response, most BC patients turn refractory to HER2-targeted drugs as early as one year after initiation of treatment. SEV-containing lncRNA AFAP1-AS1 (AFAP1 antisense RNA 1) overexpression was associated with poor prognosis in triple-negative TNBC patients where its upregulation activated Wnt/β-catenin pathway to promote tumorigenesis and cell invasion by increasing the expression of C-myc and epithelial-mesenchymal transition-related molecules [394]. It also functioned as a miR-2110 sponge to increase Sp1 expression, the AFAP1-AS1/miR-2110/Sp1 axis behaving as a potent modulator of the proliferation, migration, and invasion of triple-negative TNBC cells [395]. Another lncRNA, AGAP2-AS1, is also dysregulated in trastuzumab-resistant BC cells and plays a critical role in enhancing trastuzumab resistance by packaging into SEVs in an hnRNPA2B1-dependent manner [396]. SEV-mediated transfer of circHIPK3 also enhanced trastuzumab resistance [397]. SEV miR-1246 and miR-155 presence can be used as predictive and prognostic biomarkers for trastuzumab-based therapy resistance in HER2-positive BC [398]. Aside from the direct effect on cell proliferation, SEVs can enhance resistance to the anti-tumor immune response. SEVs from HER2-resistant cells have increased amounts of the immunosuppressive cytokine TGFβ1 and the lymphocyte activation inhibitor PD-L1, suggesting that they can induce immune evasion through neuromedin U [399].
CDK4/6 inhibition is now part of the array of targeted tools for patients with ER+ BC. SEVs’ miR-432-5p levels were higher in CDK4/6 resistant patients. Increased CDK6 expression is commonly observed in resistant cells and depends on TGF-b pathway suppression via miR-432-5p expression [400]. High baseline CDK4 mRNA levels in SEVs have been associated with response to palbociclib plus hormonal therapy, while the increase in TK1 and CDK9 mRNA copies/mL is associated with clinical resistance [401]. Deep proteomic analysis of plasma SEVs from resistant patients will help better understand underlying resistance mechanisms and give new potential resistance biomarkers [402].
To the best of our knowledge, no reports involving SEVs in the resistance process of BC cells have been yet reported for either other kinases, PARP, PI3K, mTOR, or immune checkpoint inhibitors.

5. Exosomes as Relevant Breast Cancer Biological Markers

Diagnosing BC in the early stages can make an essential difference in the patient’s treatment and prognosis. Aside from breast imaging which is crucial in the screening, diagnosis, and preoperative work-up of BC, biomarkers can provide additional insight into a patient’s diagnosis, prognosis, and response to treatment. Numerous biomarkers are currently used in BC management, notably tissue marker expression of different receptors (estrogen receptor (ER), progesterone receptor (PR), and human epidermal growth factor receptor 2 (HER2)) that is daily used for patients’ staging [403,404]. Aside from tissue ones, blood biomarkers are attractive means to monitor disease recurrence or progression, to follow treatment response, or to evidence targetable mutations that will direct therapy. Nevertheless, while iterative measurements of serum proteins such as CA15-3, CA27-29, and CEA have proved valuable tools in advanced cancers to monitor BC response to treatment, their poor sensitivity in early BC impairs their use in either diagnosis or prognosis [405]. New predictive and prognostic protein markers are still mandatory [406]. Liquid biopsies, which are supposed to get tumor-derived materials such as tumor DNA, RNA, and intact tumor cells in body fluids, are less invasive than tissue biopsies. They appear as an alternative for discovering new biomarkers for BC screening to diagnosis, prognosis, treatment response, and discovery of relapse [407,408]. As part of liquid biopsy, SEVs can be detected in patients’ biological fluids, such as blood, urine, CSF, and saliva [38] where they remain stable and protected from the degradation of serum ribonucleases and DNases [409]. SEVs can now be easily isolated [410] even though a universal standardized and widely accepted method for isolating and then analyzing SEVs is still mandatory [411,412,413,414]. As several miRNAs, lncRNAs, and proteins are differentially expressed in SEVs originating from tumor and normal cells, they are likely to be potential sources of biomarkers and become a promising field in BC management (Figure 4).
As valuable BC biomarker sources, SEVs can be divided into surface protein biomarkers and intraluminal biomarkers (mostly nucleic acids, among which figure miRNAs).

5.1. Exosome Nucleic Acid Cargo as Biomarkers

5.1.1. Exosome mRNAs as Interesting BC Markers

Messenger RNAs (mRNAs) encapsulated within SEVs are transferred to recipient cells and translated into proteins, altering the behavior of the cells [166,415,416]. Highly cancerous cells communicated with less cancerous cells through SEVs transfer, increasing migratory behavior and metastatic capacity [417]. Fragile mRNAs encapsulated within the SEVs’ phospholipid bilayer structure are protected from the harsh external environment, which would otherwise degrade them [81,418]. While SEVs transcriptomic profile reflects only partly that of the cell of origin [419], the whole transcriptomic analysis identifies a global SEV mRNA signature and BC signal in patients [420]. A typical “stemness and metastatic” signature was reported in SEVs of patients with worse prognosis. This signature comprises several mRNAs, such as those coding for NANOG, NEUROD1, HTR7, KISS1R, and HOXC6 [217].

5.1.2. SEVs miRNAs as Relevant BC Biological Markers

Circulating miRNAs (c-miRNAs) can travel in the bloodstream in two forms: in cell-free miRNAs (Ago2-related) or embedded in circulating tumor cells (CTCs), apoptotic bodies, or SEVs [421]. SEVs miRNAs are stable as they are protected from serum RNases [422]. Many papers have been published about using c-miRNAs in BC (for review [423]). A vast number of those report their modified expression either being up-or downregulated (for extensive reviews, [424,425,426,427]).
Either single miRNA (miR-21 [428], miR-155 [429], miR-223-3p [430], mir-373 [431], and mir-7641 [432]) or a combination of two or multiple miRNAs have been reported. Indeed, numerous reports have designed specific miRNA sets associated with BC. Combination of two plasma SEVs miRNAs (miR-21 and miR-1246 [433], miR-21 and miR-221 [434], miR-21 and miR-155 [435], and miR-92a, and miR-25-3p [436]), three miRNAs (miR-16, miR-30b, and miR-93 [437]; miR-21, miR-105, and miR-222 [438]; miR-21, miR-155, and miR-365 [439]; and miR-145, miR-155, and miR-382 [440]), four miRNAs (miR-21, miR-55, miR-10b, and Let-7a [441]), up to thirteen miRNAs (miR-21-3p, miR-192-5p, miR-221-3p, miR-451a, miR-574-5p, miR-1273g-3p, miR-152, miR-22-3p, miR-222-3p, miR-30a-5p, miR-30e-5p, miR-324-3p, and miR-382-5p) [442] have been described. To optimize and find new relevant miRNA combinations, some algorithms have been developed to detect BC specifically ([443].
As diagnosing BC at an early stage is still challenging, several sets of c-miRNAs have been specifically assayed for that purpose. Both sensitivity and specificity of SEVs miR-17-5p concentration were superior to conventional serum biomarkers such as CEA and CA15-3 [444]. A combination of five miRNA, miR-1246, miR-1307-3p, miR-4634, miR-6861-5p, and miR-6875-5p, was shown to detect BC with high sensitivity, specificity, and accuracy, even in the case of ductal carcinoma in situ (DCIS) [445]. Both increased concentrations of miR-21-5p and miR-10b-5p levels in serum-derived SEVs of BC patients correlate with BC grade [446]. The overall expression of nine microRNAs was higher in patients with stages I, II, and III compared to stage IV, with potential utilization for early detection [447]. Serum miR-423-5p was significantly associated with the tumor’s clinical stage and Ki-67 level [448]. A combination of miR-375, miR-655-3p, miR-548b-5p, and miR-24-2-5p has been found relevant for early BC diagnosis [449]. A dual microRNA signature based on miR-30b-5p and miR-99a-5p levels in plasma is a good diagnostic biomarker for BC [450]. Combinations of four miRNAs (miR-1246, miR-206, miR-24, and miR-373) were reported to have a sensitivity of 98%, a specificity of 96%, and an accuracy of 97% for BC detection [451]. Very recently, the miR-15a, miR-16, and miR-221 combination turned out to be promising for BC diagnostic [452].
BC prognosis is also an important issue. A recent review reported that 110 aberrantly expressed miRNAs have been associated with prognosis in BC [453]. Association of miR-126, miR-122, miR-92-1, miR-19a, and miR-29c together with circular miRNAs, such as miR-21-5p, miR-96-5p, and miR-125b-5p, can provide a promising evaluation marker in BC prognosis [454]). Moreover, a specific set of plasma SEVs miRNAs can be used for staging to evidence BC subtypes. Analysis of SEVs derived from plasma of 435 HER2+ and TNBC subtypes has identified 18 exosomal miRNAs that differed between HER2+ and TNBC subtypes, nine miRNAs also differing from healthy women [455]. Association of miR-34 and miR-520 can be used for ER+ and TNBC subtypes [454]. Such kinds of sets can also be used to predict early recurrence or metastasis. Seven miRNAs were differentially expressed between BC patients with and without recurrences, including four miRNAs upregulated (miR-21-5p, miR-375, miR-205-5p, and miR-194-5p) and three miRNAs downregulated (miR-382-5p, miR-376c-3p, and miR-411-5p) [456]. The association of miR-19a, miR-20a, miR-126, and miR-155 can discriminate against the metastatic outcome of BC patients [457,458].
A set of dysregulated selected exosomal miRNAs that could modulate target genes responsible for MAPK, TGF-beta, Wnt, mTOR, and PI3K/Akt signaling pathways have been associated with DOX resistance [384]. A specific miRNA signature was differentially expressed in SEVs derived from adriamycin-resistant (A/exo) and parental breast cancer cells (S/exo), with 309 miRNAs being increased and 66 significantly decreased in A/exo compared with S/exo [383]. The association of miRNAs targeting metabolic pathways has been reported with differential response to neoadjuvant chemotherapy (NACT) [459]. Three miRNAs before NACT (miR-30b, miR-328, and miR-423) predicted complete pathological response (pCR) in BC while upregulation of miR-127 correlated with pCR in triple-negative TNBC patients [460]. In a meta-analysis review, 60 of 123 reported miRNAs in the literature were found to be related to NACT response [461]. Dynamic evaluation of three miRNAs, including miR-222, miR-20a, and miR-451 was associated with NACT chemo-sensitivity [462]. Interestingly, a combined signature of four miRNAs (miR-4448, miR-2392, miR-2467-3p, and miR-4800-3p) could be used to discriminate between chemotherapy responders and nonresponders TNBC patients [463].
Given the vast number of publications on miRNA differential expression in BC patients’ SEVs, the development of relevant meta-analysis is strongly mandatory, and several have so far been performed. One suggested that miR-21 is likely to be a potential biomarker for early diagnosis, with high sensitivity and specificity being significantly upregulated in BC [464]. This result was confirmed later [465]. Another reported that plasma SEVs miR-23b upregulation is linked to poor overall BC survival [466]. A third one pinpoints miR-9 as an interesting BC biomarker [467]. Globally, there is little consistency among the circulating miRNA signatures identified in these different studies, mainly due to the lack of standardization and result reproducibility, which remains the most significant issue [468]. So far, no panels of circulating miRNAs are still ready for BC diagnosis in a clinical setting [469,470].

5.1.3. SEVs lncRNAs as Interesting Emerging Biomarkers

Long-noncoding RNAs are regulatory transcripts longer than 200 nucleotides that play an essential part in many fundamental cellular processes [471], and their deregulation is considered to contribute to carcinogenesis [472] and metastasis [473,474]. The increased serum concentration of several SEVs lncRNAs has been associated with poor prognosis in BC patients. LncRNA DANCR [475] and lncRNA metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) [476] have been associated with BC worsened evolution. Serum exosomal lncRNA XIST has been described as a potential biomarker to diagnose TNBC recurrence [477]. Fifteen exosome-related differentially expressed lncRNAs were recently identified to be correlated with BC prognosis [478].
Overexpression of specific lncRNAs has been evidenced as a marker of treatment resistance. Trastuzumab resistance is associated with the action of LncRNA OIP5-AS1 through miR-381-3p/HMGB3 axis [479], lncRNA ATB by competitively binding miR-200c, upregulating ZEB1 and ZNF-217, then inducing EMT [480], lncRNA AGAP2-AS1 by inducing BC cells autophagy [481], and lncRNA SNHG14 [482]. Higher expression levels of exosomal lncRNA-H19 compared to parental cells have been reported in DOX resistance [380]. A recent meta-analysis has confirmed that lncRNAs in SEVs could be a promising bioindicator for the diagnosis and prognosis of solid tumors [483].

5.1.4. SEVs Circular Nucleic Acids as New Potential Diagnostic Tool

Extrachromosomal circular DNA (circDNA) is a type of cell-free DNA (cfDNA) that is more structurally stable than linear cfDNA currently used for cancer-related detections in clinical settings [484]. CirDNA is resistant to the action of extracellular nucleases due to the formation of macromolecular complexes with proteins (including histones) [485]. Commonly observed in both standard and cancer cells (Wang 2021), it can bind to the outer surface of exosomes (Tamkovich 2016) (Tutanov 2022) and be detected in serum (Ling 2021).
Aside from circDNA, circular RNAs (circRNA) also exist. CircRNA is a class of covalently closed single-stranded circular RNA molecules without free 5′ or 3′ ends [486]. Unlike traditional linear RNAs such as lncRNAs and miRNAs, circRNAs were not degraded by RNases or RNA exonucleases and were more stable and conserved in peripheral blood or plasma [487]. CircRNAs can exert various functions according to their parental genes, among which figure their ability to serve as a sponge for multiple miRNAs, suppressing their activity [488]. Many circRNAs have been discovered in various cancers, and they are activated in either inhibiting tumor progression or promoting tumorigenesis.
Both cDNA and circRNA can be observed in BC and hold promise to be used as SEVs biomarkers. CirRNA circ_0004771 accelerates BC cell carcinogenic phenotypes via upregulating dimethylarginine dimethylaminohydrolase 1 (DDAH1) expression through absorbing miR-1253 [489]. Circ_0000615, which was spliced from the ZNF609 gene, displays an expression level markedly upregulated in BC cell lines compared with normal ductal epithelial cells. It displays a better diagnostic efficiency in BC patients than routine tumor biomarkers such as CA153, CA125, and CEA. Its high expression was closely associated with advanced tumor stage, lymph node metastasis, and high grade of recurrence risk [490]. In TNBC, several cirRNAs are likely to be exciting biomarkers. lncRNA MALAT1 (metastasis-associated lung adenocarcinoma transcript 1) regulates linear isoforms of VEGFA, inducing the back-splicing of VEGFA exon 7 and producing circular RNA circ_0076611. Circ_0076611 is detectable in TNBC cells [491]. The expression of circHSDL2, targeting let-7a-2-3p during the progression of TNBC, was found to be significantly upregulated in serum SEVs and tumor tissues from TNBC patients [492]. Overexpression of circ-proteasome 20S subunit alpha 1 (circ-PSMA1) promoted tumorigenesis, metastasis, and migration through miR-637/Akt1/β-catenin (cyclin D1) axis in TNBC both in vitro and in vivo. circ-PSMA1 is upregulated in vitro in TNBC cells’ SEVs and in SEVs isolated from triple-negative TNBC patients’ sera [493].

5.2. SEVs Protein Cargo as a Source of New Cancer Biomarkers

Proteins located on the surface of and within SEVs may also be used as relevant cancer biomarkers as they may differ between healthy individuals and BC patients [494]. SEVs surface protein markers such as members of the tetraspanin family (CD9, CD63, CD81, CD82, and CD151), some integrins, multivesicular bodies (MVBs) formation proteins (TSG101, Alix, and Clathrin), and lipid raft proteins (flotillins) [495]. The level of CD82 was significantly higher in the serum of BC patients compared to the healthy controls, while the expression of CD82 significantly increased with malignant BC progression [496]. On the contrary, CD151-deleted SEVs significantly decreased the migration and invasion of TNBC cells [497]. In addition to these self-proteins for constructing SEVs, several proteins from BC-derived cells are likely to be potential biomarkers for early screening and diagnosis.
Enzymes and specific signaling proteins (EpCAM, EFGR, and survivin-2B) along with metalloproteinase ADAM10, heat-shock protein HSP70, and Annexin-1 can also be evidenced as general marker proteins detected in serum and pleural effusion-derived SEVs from BC cell lines or BC patients [237,498]. Epithelial cell adhesion molecules such as EpCAM and CD24 could be used as markers to identify cancer-derived SEVs in ascites and pleural effusions from BC patients [498]. As compared to healthy controls, higher levels of SEVs with glypican 1 (GPC1) on their surface (GPC1+) are found in BC patients’ sera (Melo 2015). GPC-1, glucose transporter 1 (GLUT-1), and disintegrin ADAM10 were potential TNBC biomarkers [237]. FAK and EGFR proteins can also be found, where FAK presence in SEV fractions is associated with in situ and stages I–III, while EGFR is associated with in situ and stage I BC [499,500]. SEVs containing amphiregulin (AREG) which binds to cell surface EGFR, were revealed to increase receptor BC invasive ability of cells [501] and can be used as prognostic and/or predictive markers [502]. Comparative proteome analysis of circulating SEVs in healthy and BC patients has shown that the association of three favorable (Serpin A1 (SERPINA1), keratin 6 (KRT6B), and SOCS3) and one unfavorable (insulin growth factor 2 receptor (IGF2R)) SEV protein markers allow diagnosing with 73% sensitivity and 100% specificity BC stage I and II [495]. The diagnostic value of fibronectin [503] and developmental endothelial locus-1 (Del-1) [504] in BC cell-derived SEVs were reported to display a sensitivity of 94.70% and a specificity of 86.36%. Some other markers are promising. Serum SEV annexin2 (AnxA2) holds promise as a potential prognosticator of TNBC as it is high in African American women with TNBC (Chaudhary 2020). The distinct expression pattern of SEV survivin-2B in serum is considered a sign of early-stage BC [505]. Some specific SEVs proteins have been correlated to response to treatment. Enrichment of CD44 in SEVs of doxycycline DOX-treated BC cells promotes their chemoresistance [506]. Programmed death ligand-1 (PD-L1), an essential immune checkpoint molecule, is expressed on BC SEVs and correlated with the progression and immunotherapy response [507]. Plasma SEV NGF concentration in BC patients undergoing neoadjuvant chemotherapy is associated with significantly poorer overall survival [508].

6. SEVs as Attractive Targets to Inhibit BC

SEVs are a source of cancer dissemination and a promoter of patients’ resistance to treatment. It is, therefore, mandatory to explore new therapeutic possibilities to suppress SEV-induced tumor progression and reduce SEV-related drug resistance.

6.1. Inhibition of SEV Uptake by Target Cells

The first possibility to limit SEVs’ adverse effects would be to inhibit their uptake by target cells [509]. SEV uptake capability has been reported to vary depending on the recipient cell type but not on the donor cell type [109]. It largely depends on surface molecules and glycoproteins on the vesicle membrane and the plasma membrane of the recipient cell [510]. Multiple uptake mechanisms are involved in the cellular internalization of SEVs, including caveolin- or clathrin-dependent endocytosis, macropinocytosis, phagocytosis, lipid raft-mediated internalization, and membrane fusion [510,511]. Many studies have found pharmacological inhibitors that could inhibit SEVs internalization. Heparin can inhibit SEVs uptake in a dose-dependent manner through direct action on heparan sulfate proteoglycans which themselves play a role in SEVs endocytosis [512]. Both cytochalasin D, through a direct inhibitory effect on actin polymerization, and methyl-β-cyclodextrin (MβCD), by depleting membranes’ cholesterol hence disrupting lipid rafts stability, inhibit phagocytosis/endocytosis mechanisms, and thus SEV uptake [513,514]. Disruption of clathrin-mediated and caveolin-dependent endocytosis by chlorpromazine or dynasore, a specific inhibitor of dynamin 2, as well as macropinocytosis inhibition by amiloride or omeprazole (OME) also inhibits SEVs uptake [515,516,517,518]. However, the extensive repertoire of mechanisms involved in SEV uptake in cancer impairs the overall efficiency of these molecules.

6.2. Inhibition of SEV Biogenesis

Another way to limit SEVs action would be to inhibit SEV biogenesis. Such an issue involves complex mechanisms and is likely to be challenging to implement. However, many pharmacological agents have been found and seem promising. The fluidity of the cell plasma membrane is fundamental during membrane lipid bilayer re-organization and SEV formation. In cancer, lipid mediators such as sphingosine 1-phosphate and ceramide, which are known to be associated with inflammation [519], also regulate SEV production [32,520]. Sphingomyelinases, acid (SMPD1), and neutral sphingomyelinase (SMPD2) are ubiquitous enzymes required for ceramide synthesis that can be specifically inhibited. GW4869, cambinol, and spiroepoxide inhibit SMPD2. Blocking SMPD2 by either GW4869 drug or specific SMPD2 siRNA results in a dose-dependent inhibition of SEV release [521]. GW4869 blocks SEV biogenesis by preventing the ceramide-modulated inward budding of multivesicular bodies and the subsequent release of SEVs [522]. Complementarily, SMPD2 overexpression increases miRNAs’ extracellular amounts [523]. The link between SMPD2 and SEVs has been associated with BC aggressiveness [253]. The association of OME that inhibits SEVs uptake with GW4869 that limits SEV biogenesis reduces paclitaxel (PTX) amount in SEVs, thus increasing the therapeutic effect of PTX on BC cells [524]. As GW4869 seems promising, imipramine, a tricyclic antidepressant, is also a source of interest because of its inhibitory activity on SMPD1 [525,526].
TSG101 is a protein involved in endosome trafficking and SEV biogenesis [527]. TSG101 knockdown in BC cells induces apoptosis and inhibits proliferation, suggesting that TSG101 is a potential therapeutic target in cancer [528].

SEV Release Inhibition

A third possibility to target SEVs relies on limiting or inhibiting their release. A drug that can inhibit SEV release is manumycin A, an antibiotic that is a selective and robust inhibitor of Ras farnesyl transferases. Farnesyltransferase inhibitors inhibit Ras activity and, therefore, SEV release [522]. Rasal2, a Ras-GTPase-activating protein (RasGAP), is a known tumor suppressor in luminal B breast cancer, frequently metastatic and recurrent. Rasal2 knockout (KO) in MCF-7 cells enhanced SEVs release and increased autophagy-related proteins in exosomal fraction while attenuated by SEV release inhibitor GW4869 (Wang 2019). Aside from Ras proteins, there are also Rab proteins that are also modulators of SEVs biogenesis [12].
Interestingly, associated with an increase in SEV secretion, the most up-regulated proteins in long-term estrogen-deprived MCF-7 LTED cells were represented by Rab GTPases [363]. Among Rab proteins, Rab27a and Rab27b seem to play a significant role in SEVs docking and exocytosis [48] and are involved in mammary gland development [125] and cancer [529]. Either knockdown of Rab27a in lung cancer [530] or gold nanoparticles conjugated with anti-sense Rab27a oligonucleotides to mute Rab27a in BC [531] generate significant inhibition of SEV release.
As plasma membrane fluidity is essential for SEV shedding, drugs targeting lipid raft formation or cholesterol synthesis will interfere with SEV release. Lipid depletion results in SEV release reduction (Skotland 2017). Pantethine, a pantothenic acid (vitamin B5) derivative, is used as an intermediate in the production of coenzyme A and plays a role in lipid metabolism, reducing total cholesterol levels. Pantethine inhibits cholesterol synthesis by 80% and fatty acid synthesis (Ranganathan 1982). Pantethine prevents murine systemic sclerosis by inhibiting microparticle shedding (Kavian 2015), an effect also observed on chemoresistant BC cells (Roseblade 2015).
Actin and actin-regulating proteins are also strongly involved in SEV secretion. Invadopodia are cellular structures used by cancer cells to degrade extracellular matrix and invade. Because of high levels of actin, such structures are critical sites for EV release. Indeed, invadopodia inhibition limits EV release [532]. Targeting cortactin, the actin-nucleation-promoting factor acting as an actin dynamics regulator, decreased SEV release, whereas its overexpression increased [533]. The non-receptor tyrosine kinase Pyk2 is highly expressed in BC and mediates invadopodia formation and function via interaction with cortactin. Targeting Pyk2 with a specific Pyk2-derived peptide inhibits invadopodia-mediated breast cancer metastasis [534].
Other drugs targeting SEV release have been used in BC. A novel anti-cancer SMR peptide that antagonizes BC cell SEVs release results in cell cycle arrest and tumor growth suppression [535]. D-Rhamnose β-hederin (DRβ-H), a novel oleanane-type triterpenoid saponin [536], and shikonin, a naphthoquinone [537], both isolated from traditional Chinese medicinal plants, attenuate resistance traits in doxycycline DOX-resistant BC cells and reduce tumor burden by decreasing SEV secretion. Cannabidiol (CBD) has been reported to be a potential inhibitor for SEV release in BC as it inhibits, in a dose-dependent manner, SEV release in MDA-MB-231 cells [538]. PEG-SMRwt-Clu, a drug derived from the secretion region of HIV-1 Nef protein, regulates exosomal pathway trafficking and seems promising. PEG-SMRwt-Clu was able to inhibit cell growth in BC cell lines and, more interestingly, to increase chemosensitivity partially. PEG-SMRwt-Clu was also associated with a decrease in the number of released SEVs [539].
Despite the current efforts and the number of SEV endocytosis, biogenesis, and release inhibitors already available, SEV inhibition remains a very complex issue because of the multifactorial nature of the different pathways involved in these processes. Nevertheless, there is no doubt that SEV uptake, biogenesis, or release inhibition is still a potential and attractive therapeutic cancer target.

7. SEVs as Nanovectors to Drive Therapy in BC

SEVs are significant players in tumor progression via the transfer of the cargo within them. Another possible way to cure BC would be to use an SEV-based therapy that uses SEVs as therapeutic nanovectors.
In the very last years, several reports have mainly focused on the idea that SEVs could be natural delivery vehicles to transport therapeutic drugs, antibodies, or RNAs to modify gene expression, especially in the cancer field [540,541,542,543,544,545,546] with a specific dedication to BC [547,548,549,550]. Indeed, SEVs are biocompatible, biodegradable, and, therefore, less toxic and immunogenic than other nanoparticle drug delivery systems such as liposomes or polymeric nanoparticles [551]. SEVs have innate limited immunogenicity and cytotoxicity [552] and can pass through anatomical barriers [553]. Additionally, as SEVs avoid drug degradation by extracellular enzymes, drug stability is enhanced [554]. Altogether, SEV’s capacity to target tumor cells is ten times higher than liposomes of a similar size. Such property is undoubtedly linked to particular ligand-receptor interactions and to efficient endocytosis mechanisms linked to the SEV membrane lipid composition that contributes significantly to cellular adherence and internalization [555].
Several reports have demonstrated the potential of using SEV therapy, and clinical trials are currently underway to find the best treatments that extend patient survival. Many kinds of SEV-based therapies have been shown to improve chemotherapy effectiveness. SEVs have been used to deliver chemotherapeutic drugs such as paclitaxel (PTX) [556,557,558] or doxycycline (DOX) [559,560,561]. While loading DOX in SEVs reduces its cardiotoxicity [562,563], it also enhances its efficacy when compared to traditional administration [562,564]. Packaging DOX into SEVs increases its stability, thus allowing a better collection within the tumor [564] with more limited side toxicity [565]. It holds the same for PTX, SEVs being more efficient than free PTX and liposomal PTX in inhibiting cancer cell growth [566]. However, developing SEV fusion with liposomes to produce a hybrid exosome (HE) with improved PTX loading capacity and enhanced tumor-targeting ability seems promising for triple-negative TNBC chemotherapy [567]. Loaded SEVs can overcome drug efflux transporter adverse effects, decreasing tumor metastasis compared to controls [568]. Interestingly, SEVs can provide cargo combinational therapy, as shown for the PTX/5-FU association in BC [558]. Very recently, lapatinib-loaded exosomes were developed as a drug delivery system in BC (Değirmenci 2022).
Aside from drug transport, SEVs are natural nucleic acid molecule carriers and can be genetically engineered to deliver specific DNA or RNA molecules. More recently, exosome–liposome hybrid nanoparticles have been developed to deliver the gene editing system CRISPR/Cas9 in mesenchymal stromal cells (MSCs) [569]. SEV vectorization of specific miRNAs has also been used. BC cell proliferation and migration were significantly suppressed when cells were treated with SEVs loaded with miR-142-3p [570] and let7c-5p [571]. EGFR-expressing cells can be targeted with GE11-positive SEVs loaded with miR-let-7a, a tumor suppressor microRNA. The results showed an efficient delivery of SEV cargo and tumor growth inhibition [572]. While miR-134 SEV delivery has been shown to enhance TNBC cells’ drug sensitivity [573], TNBC aggressiveness was suppressed using either miR-381-3p- or miR-145-containing MSC-derived SEVs [574,575]. A synergistic efficacy of co-delivering miR-159 and DOX in SEVs was reported for TNBC therapy [576]. Not only miRNAs can be transferred through SEV delivery. DARS-AS1, a newly reported CUMS-responsive lncRNA, was enriched in TNBC cells and positively correlated with the late clinical stage in patients with TNBC. Treatment with DARS-AS1 siRNA-loaded SEVs substantially slowed CUMS-induced TNBC cell growth and liver metastasis [577].
SEVs can also be used as a new type of tumor vaccine [578]. SEVs have been explored as modulators of the immune response against tumor cells. In BC, treatment with topotecan (TPT, an inhibitor of topoisomerase I) induces BC cells to release SEVs containing DNA that activates dendritic cells (DC) [579]. DCs are antigen-presenting cells that are central to the initiation and regulation of innate and adaptive immunity in the tumor microenvironment. DCs have been shown to secrete antigen-presenting SEVs that coexpress major histocompatibility complex molecules. Such SEVs activate specific cytotoxic T lymphocytes in vivo that can reduce or SEVen suppress tumor growth [580]. SEVs from DCs are likely to initiate an immune response against tumor cells more precisely and accurately than cell therapy and other non-cell-based therapy [581]. Vaccination of transgenic HLA-A2/HER2 mice with a single dose of SEVs from DCs transfected with an adenoviral vector led to activating CD8+ T cell cytolytic functions against BC cells in vitro and reduced tumor growth in vivo [582].
Interestingly, tumor cell-derived SEVs have been shown to have an immunostimulatory effect on antitumor DCs [583]. Such evidence prompts to start engineering DC using targeted-SEV delivery of antigens and adjuvants to DCs, representing a fundamental approach for developing DC vaccines [584]. BC cell line 4T1-derived SEV-mediated transfer of let-7i, miR-155, and miR-142 to DC enhances DC maturation [585].
Cells under different conditions will determine SEV heterogeneity, generating vast and complex combinatorial possibilities. Cell-derived SEVs are generally directed to specific cell types [12]. SEVs derived from hypoxic tumor cells tend to be more easily taken up by hypoxic tumor cells [586]. Thus, to better use SEVs in cancer, engineering SEVs with ligands that can specifically bind to targeted cancer cells is mandatory. Either SEV surface expression of receptor/ligand, antibody/ligand, or microenvironment-specific molecules can be used to modify SEVs. Recently, bioengineered SEVs have been able to specifically bind to HER2 by expressing designed ankyrin repeat proteins (DARPins) on their membrane surface [587]. Directing CD3 and EGFR expressions on SEV membranes was shown to induce cross-linking of T cells and EGFR-expressing BC cells and elicit potent antitumor immunity both in vitro and in vivo [588]. It holds the same when SEVs are engineered through the genetic display of anti-human CD3 and anti-human HER2 antibodies, dually targeting CD3 T cell and BC-associated HER2 receptors. Such SEVs redirect and activate cytotoxic T cells toward attacking HER2-expressing BC cells [589]. Both hyaluronan (HA), the most specific CD44 ligand [590], and CD44 itself, which are both mainly involved in the metastasis process, have been evidenced in BC cells EVs and associated with chemoresistance [506]. High accumulation of HA in the tumor microenvironment leads to an increase in the interstitial pressure and reduced perfusion of drugs. Hyaluronidase, an enzyme that degrades HA, has been engineered into SEV. Hyaluronidase-containing SEVs have been developed and shown to degrade tumor extracellular matrix and enhance the permeability of T cells and drugs within the tumor [591], inhibiting BC metastasis and improving tumor treatment efficiency [592]. Another smart reported strategy was to use HA-engineered SEVs to direct chemotherapy to CD44 expressing BC cells. HA decoration of milk DOX-containing SEVs directs tumor-specific delivery of DOX [593].
Using SEVs as therapeutic vectors in cancer seems very promising, and clinical trials are nowadays being carried out [594]. Unfortunately, breakthroughs still need to occur because of the complexity of handling such new therapeutic methods in vivo. There is also an urgent need to better understand SEV biology and nature to accelerate SEV vectorization in BC patient treatment.

8. Conclusions

It is now clearly stated that SEVs exert various biological functions, mainly via delivering signaling molecules that regulate an extensive repertoire of cellular processes. Their role in cancer development seems central as they are significant players in multidirectional signaling between cancer cells and various other ones (from neighboring tumor microenvironment cells at the primary tumor site to more distant ones). It covers every step of BC carcinogenesis up to metastatic dissemination. SEV detection in a large variety of biological fluids represents the future of cancer detection, an easy and reproducible means to identify new specific diagnostic and prognostic biomarkers. SEVs also represent new targets for treatment as their inhibition could limit or stop cancer development. Additionally, these extracellular signaling cargos could be used as specific vectors to convey conventional or innovative therapies to targeted cancer cells.
However, fundamental research is still mandatory to understand SEV function in cancer progression. Although pre-clinical data appear very promising, validation from large clinical trials is needed to support the daily use of SEVs as either tumor biomarkers for monitoring cancer progression and driving treatment decisions or new vectors for specifically targeted treatments.

Author Contributions

S.L. and J.A.D. wrote the paper; C.D., M.S. and M.C. extensively reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  2. Ferlay, J.; Colombet, M.; Soerjomataram, I.; Dyba, T.; Randi, G.; Bettio, M.; Gavin, A.; Visser, O.; Bray, F. Cancer Incidence and Mortality Patterns in Europe: Estimates for 40 Countries and 25 Major Cancers in 2018. Eur. J. Cancer 2018, 103, 356–387. [Google Scholar] [CrossRef]
  3. Gennari, A.; André, F.; Barrios, C.H.; Cortés, J.; de Azambuja, E.; DeMichele, A.; Dent, R.; Fenlon, D.; Gligorov, J.; Hurvitz, S.A.; et al. ESMO Clinical Practice Guideline for the Diagnosis, Staging and Treatment of Patients with Metastatic Breast Cancer. Ann. Oncol. 2021, 32, 1475–1495. [Google Scholar] [CrossRef] [PubMed]
  4. Harbeck, N.; Penault-Llorca, F.; Cortes, J.; Gnant, M.; Houssami, N.; Poortmans, P.; Ruddy, K.; Tsang, J.; Cardoso, F. Breast Cancer. Nat. Rev. Dis. Primers 2019, 5, 66. [Google Scholar] [CrossRef]
  5. Luque-Bolivar, A.; Pérez-Mora, E.; Villegas, V.E.; Rondón-Lagos, M. Resistance and Overcoming Resistance in Breast Cancer. Breast Cancer Targets Ther. 2020, 12, 211–229. [Google Scholar] [CrossRef] [PubMed]
  6. Saha, T.; Lukong, K.E. Breast Cancer Stem-Like Cells in Drug Resistance: A Review of Mechanisms and Novel Therapeutic Strategies to Overcome Drug Resistance. Front. Oncol. 2022, 12, 856974. [Google Scholar] [CrossRef]
  7. Calaf, G.M.; Zepeda, A.B.; Castillo, R.L.; Figueroa, C.A.; Arias, C.; Figueroa, E.; Farías, J.G. Molecular Aspects of Breast Cancer Resistance to Drugs (Review). Int. J. Oncol. 2015, 47, 437–445. [Google Scholar] [CrossRef] [Green Version]
  8. Nazemi, M.; Rainero, E. Cross-Talk Between the Tumor Microenvironment, Extracellular Matrix, and Cell Metabolism in Cancer. Front. Oncol. 2020, 10, 239. [Google Scholar] [CrossRef] [Green Version]
  9. Pernot, S.; Evrard, S.; Khatib, A.-M. The Give-and-Take Interaction Between the Tumor Microenvironment and Immune Cells Regulating Tumor Progression and Repression. Front. Immunol. 2022, 13, 850856. [Google Scholar] [CrossRef] [PubMed]
  10. Cox, T.R. The Matrix in Cancer. Nat. Rev. Cancer 2021, 21, 217–238. [Google Scholar] [CrossRef]
  11. Naito, Y.; Yoshioka, Y.; Yamamoto, Y.; Ochiya, T. How Cancer Cells Dictate Their Microenvironment: Present Roles of Extracellular Vesicles. Cell Mol. Life Sci. 2016, 74, 697–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Colombo, M.; Raposo, G.; Théry, 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]
  13. Margolis, L.; Sadovsky, Y. The Biology of Extracellular Vesicles: The Known Unknowns. PLoS Biol. 2019, 17, e3000363. [Google Scholar] [CrossRef] [PubMed]
  14. Shah, R.; Patel, T.; Freedman, J.E. Circulating Extracellular Vesicles in Human Disease. N. Engl. J. Med. 2018, 379, 958–966. [Google Scholar] [CrossRef] [PubMed]
  15. Xu, R.; Rai, A.; Chen, M.; Suwakulsiri, W.; Greening, D.W.; Simpson, R.J. Extracellular Vesicles in Cancer—Implications for Future Improvements in Cancer Care. Nat. Rev. Clin. Oncol. 2018, 15, 617–638. [Google Scholar] [CrossRef]
  16. Latifkar, A.; Cerione, R.A.; Antonyak, M.A. Probing the Mechanisms of Extracellular Vesicle Biogenesis and Function in Cancer. Biochem. Soc. Trans. 2018, 46, 1137–1146. [Google Scholar] [CrossRef]
  17. Rahbarghazi, R.; Jabbari, N.; Sani, N.A.; Asghari, R.; Salimi, L.; Kalashani, S.A.; Feghhi, M.; Etemadi, T.; Akbariazar, E.; Mahmoudi, M.; et al. Tumor-Derived Extracellular Vesicles: Reliable Tools for Cancer Diagnosis and Clinical Applications. Cell Commun. Signal. 2019, 17, 73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Nazarenko, I. Recent Developments in Technology and Perspectives for Cancer Liquid Biopsy. In Tumor Liquid Biopsies; Recent Results in Cancer Research; Springer: Cham, Switzerland, 2020; Volume 215, pp. 319–344. [Google Scholar] [CrossRef]
  19. Liu, J.; Chen, Y.; Pei, F.; Zeng, C.; Yao, Y.; Liao, W.; Zhao, Z. Extracellular Vesicles in Liquid Biopsies: Potential for Disease Diagnosis. Biomed. Res. Int. 2021, 2021, 6611244. [Google Scholar] [CrossRef]
  20. Théry, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal Information for Studies of Extracellular Vesicles 2018 (MISEV2018): A Position Statement of the International Society for Extracellular Vesicles and Update of the MISEV2014 Guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef] [Green Version]
  21. Witwer, K.W.; Théry, C. Extracellular Vesicles or Exosomes? On Primacy, Precision, and Popularity Influencing a Choice of Nomenclature. J. Extracell. Vesicles 2019, 8, 1648167. [Google Scholar] [CrossRef]
  22. Caruso, S.; Poon, I.K.H. Apoptotic Cell-Derived Extracellular Vesicles: More Than Just Debris. Front. Immunol. 2018, 9, 1486. [Google Scholar] [CrossRef] [Green Version]
  23. Cocucci, E.; Meldolesi, J. Ectosomes and Exosomes: Shedding the Confusion between Extracellular Vesicles. Trends Cell Biol. 2015, 25, 364–372. [Google Scholar] [CrossRef]
  24. Doyle, L.M.; Wang, M.Z. Overview of Extracellular Vesicles, Their Origin, Composition, Purpose, and Methods for Exosome Isolation and Analysis. Cells 2019, 8, 727. [Google Scholar] [CrossRef] [Green Version]
  25. Alli, A.A. Extracellular Vesicles Mechanisms of Extracellular Vesicle Biogenesis, Cargo Loading, and Release. Physiology 2021, 13. [Google Scholar] [CrossRef]
  26. Sherman, C.D.; Lodha, S.; Sahoo, S. EV Cargo Sorting in Therapeutic Development for Cardiovascular Disease. Cells 2021, 10, 1500. [Google Scholar] [CrossRef]
  27. Mammes, A.; Pasquier, J.; Mammes, O.; Conti, M.; Douard, R.; Loric, S. Extracellular Vesicles: General Features and Usefulness in Diagnosis and Therapeutic Management of Colorectal Cancer. World J. Gastrointest. Oncol. 2021, 13, 1561–1598. [Google Scholar] [CrossRef]
  28. Nakano, A. The Golgi Apparatus and Its Next-Door Neighbors. Front. Cell Dev. Biol. 2022, 10, 884360. [Google Scholar] [CrossRef]
  29. Wang, S.; Thibault, G.; Ng, D.T.W. Routing Misfolded Proteins through the Multivesicular Body (MVB) Pathway Protects against Proteotoxicity. J. Biol. Chem. 2011, 286, 29376–29387. [Google Scholar] [CrossRef] [Green Version]
  30. Christ, L.; Raiborg, C.; Wenzel, E.M.; Campsteijn, C.; Stenmark, H. Cellular Functions and Molecular Mechanisms of the ESCRT Membrane-Scission Machinery. Trends Biochem. Sci. 2017, 42, 42–56. [Google Scholar] [CrossRef]
  31. Olmos, Y. The ESCRT Machinery: Remodeling, Repairing, and Sealing Membranes. Membranes 2022, 12, 633. [Google Scholar] [CrossRef]
  32. Kajimoto, T.; Okada, T.; Miya, S.; Zhang, L.; Nakamura, S. Ongoing activation of sphingosine 1-phosphate receptors mediates maturation of exosomal multivesicular endosomes. Nat. Commun. 2013, 4, 2712. [Google Scholar] [CrossRef] [Green Version]
  33. Henne, W.M.; Buchkovich, N.J.; Emr, S.D. The ESCRT Pathway. Dev. Cell 2011, 21, 77–91. [Google Scholar] [CrossRef] [Green Version]
  34. Peng, X.; Yang, L.; Ma, Y.; Li, Y.; Li, H. Focus on the Morphogenesis, Fate and the Role in Tumor Progression of Multivesicu-lar Bodies. Cell Commun. Signal. 2020, 18, 122. [Google Scholar] [CrossRef]
  35. Scott, C.C.; Gruenberg, J. Ion Flux and the Function of Endosomes and Lysosomes: PH Is Just the Start. Bioessays 2011, 33, 103–110. [Google Scholar] [CrossRef]
  36. Savina, A.; Furlán, M.; Vidal, M.; Colombo, M.I. Exosome Release Is Regulated by a Calcium-dependent Mechanism in K562 Cells. J. Biol. Chem. 2003, 278, 20083–20090. [Google Scholar] [CrossRef] [Green Version]
  37. Kumar, A.; Deep, G. Hypoxia in Tumor Microenvironment Regulates Exosome Biogenesis: Molecular Mechanisms and Translational Opportunities. Cancer Lett. 2020, 479, 23–30. [Google Scholar] [CrossRef]
  38. Raposo, G.; Stoorvogel, W. Extracellular Vesicles: Exosomes, Microvesicles, and Friends. J. Cell Biol. 2013, 200, 373–383. [Google Scholar] [CrossRef] [Green Version]
  39. Kalluri, R.; LeBleu, V.S. The Biology, Function, and Biomedical Applications of Exosomes. Science 2020, 367, 1–40. [Google Scholar] [CrossRef]
  40. Klionsky, D.J.; Eskelinen, E.-L.; Deretic, V. Autophagosomes, Phagosomes, Autolysosomes, Phagolysosomes, Autophagoly-sosomes… Wait, I’m Confused. Autophagy 2014, 10, 549–551. [Google Scholar] [CrossRef] [Green Version]
  41. Xing, H.; Tan, J.; Miao, Y.; Lv, Y.; Zhang, Q. Crosstalk between Exosomes and Autophagy: A Review of Molecular Mechanisms and Therapies. J. Cell Mol. Med. 2021, 25, 2297–2308. [Google Scholar] [CrossRef]
  42. Falguières, T.; Luyet, P.-P.; Gruenberg, J. Molecular Assemblies and Membrane Domains in Multivesicular Endosome Dynamics. Exp. Cell Res. 2009, 315, 1567–1573. [Google Scholar] [CrossRef]
  43. Hessvik, N.P.; Llorente, A. Current Knowledge on Exosome Biogenesis and Release. Cell Mol. Life Sci. 2018, 75, 193–208. [Google Scholar] [CrossRef] [Green Version]
  44. Anand, S.; Samuel, M.; Kumar, S.; Mathivanan, S. Ticket to a Bubble Ride: Cargo Sorting into Exosomes and Extracellular Vesicles. Biochim. Biophys. Acta BBA-Proteins Proteom. 2019, 1867, 140203. [Google Scholar] [CrossRef]
  45. Granger, E.; McNee, G.; Allan, V.; Woodman, P. The Role of the Cytoskeleton and Molecular Motors in Endosomal Dynamics. Semin. Cell Dev. Biol. 2014, 31, 20–29. [Google Scholar] [CrossRef]
  46. Titus, M.A. Myosin-Driven Intracellular Transport. Cold Spring Harb. Perspect. Biol. 2018, 10, a021972. [Google Scholar] [CrossRef] [Green Version]
  47. Langemeyer, L.; Fröhlich, F.; Ungermann, C. Rab GTPase Function in Endosome and Lysosome Biogenesis. Trends Cell Biol. 2018, 28, 957–970. [Google Scholar] [CrossRef]
  48. 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]
  49. Song, L.; Tang, S.; Han, X.; Jiang, Z.; Dong, L.; Liu, C.; Liang, X.; Dong, J.; Qiu, C.; Wang, Y.; et al. KIBRA Controls Exosome Secretion via Inhibiting the Proteasomal Degradation of Rab27a. Nat. Commun. 2019, 10, 1639. [Google Scholar] [CrossRef] [Green Version]
  50. Hong, W.; Lev, S. Tethering the Assembly of SNARE Complexes. Trends Cell Biol. 2014, 24, 35–43. [Google Scholar] [CrossRef]
  51. Südhof, T.C.; Rothman, J.E. Membrane Fusion: Grappling with SNARE and SM Proteins. Science 2009, 323, 474–477. [Google Scholar] [CrossRef] [Green Version]
  52. Wei, Y.; Wang, D.; Jin, F.; Bian, Z.; Li, L.; Liang, H.; Li, M.; Shi, L.; Pan, C.; Zhu, D.; et al. Pyruvate Kinase Type M2 Promotes Tumour Cell Exosome Release via Phosphorylating Synaptosome-Associated Protein 23. Nat. Commun. 2017, 8, 14041. [Google Scholar] [CrossRef] [Green Version]
  53. Xu, M.; Ji, J.; Jin, D.; Wu, Y.; Wu, T.; Lin, R.; Zhu, S.; Jiang, F.; Ji, Y.; Bao, B.; et al. The Biogenesis and Secretion of Exosomes and Multivesicular Bodies (MVBs): Intercellular Shuttles and Implications in Human Diseases. Genes Dis. 2022, in press. [Google Scholar] [CrossRef]
  54. Nabhan, J.F.; Hu, R.; Oh, R.S.; Cohen, S.N.; Lu, Q. Formation and Release of Arrestin Domain-Containing Protein 1-Mediated Microvesicles (ARMMs) at Plasma Membrane by Recruitment of TSG101 Protein. Proc. Natl. Acad. Sci. USA 2012, 109, 4146–4151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Yokoi, A.; Ochiya, T. Exosomes and Extracellular Vesicles: Rethinking the Essential Values in Cancer Biology. Semin. Cancer Biol. 2021, 74, 79–91. [Google Scholar] [CrossRef] [PubMed]
  56. Juan, T.; Fürthauer, M. Biogenesis and Function of ESCRT-Dependent Extracellular Vesicles. Semin. Cell Dev. Biol. 2018, 74, 66–77. [Google Scholar] [CrossRef] [PubMed]
  57. Andrews, A.M.; Rizzo, V. Microparticle-Induced Activation of the Vascular Endothelium Requires Caveolin-1/Caveolae. PLoS ONE 2016, 11, e0149272. [Google Scholar] [CrossRef] [Green Version]
  58. Li, S.; Lisanti, M.; Puszkin, S. Purification and Molecular Characterization of NP185, a Neuronal-Specific and Syn-apse-Enriched Clathrin Assembly Polypeptide. Bioquim. Y Patol. Clin. Bypc. Rev. Asoc. Bioquim. Argent. 1998, 62, 5–17. [Google Scholar]
  59. Sun, M.; Xue, X.; Li, L.; Xu, D.; Li, S.; Li, S.C.; Su, Q. Ectosome Biogenesis and Release Processes Observed by Using Live-Cell Dynamic Imaging in Mammalian Glial Cells. Quant. Imaging Med. Surg. 2021, 11, 4604–4616. [Google Scholar] [CrossRef] [PubMed]
  60. Hugel, B.; Martínez, M.C.; Kunzelmann, C.; Freyssinet, J.-M. Membrane Microparticles: Two Sides of the Coin. Physiology 2005, 20, 22–27. [Google Scholar] [CrossRef] [PubMed]
  61. Ratajczak, J.; Wysoczynski, M.; Hayek, F.; Janowska-Wieczorek, A.; Ratajczak, M.Z. Membrane-derived microvesicles: Im-portant and underappreciated mediators of cell-to-cell communication. Leukemia 2006, 20, 1487–1495. [Google Scholar] [CrossRef]
  62. Bevers, E.M.; Comfurius, P.; Dekkers, D.W.C.; Zwaal, R.F.A. Lipid Translocation across the Plasma Membrane of Mammalian Cells. Biochim. Biophys. Acta BBA-Mol. Cell Biol. Lipids 1999, 1439, 317–330. [Google Scholar] [CrossRef]
  63. MacKenzie, A.; Wilson, H.L.; Kiss-Toth, E.; Dower, S.K.; North, R.A.; Surprenant, A. Rapid Secretion of Interleukin-1beta by Microvesicle Shedding. Immunity 2001, 15, 825–835. [Google Scholar] [CrossRef]
  64. Mir, B.; Goettsch, C. Extracellular Vesicles as Delivery Vehicles of Specific Cellular Cargo. Cells 2020, 9, 1601. [Google Scholar] [CrossRef]
  65. Isaac, R.; Reis, F.C.G.; Ying, W.; Olefsky, J.M. Exosomes as Mediators of Intercellular Crosstalk in Metabolism. Cell Metab. 2021, 33, 1744–1762. [Google Scholar] [CrossRef] [PubMed]
  66. Freeman, D.W.; Hooten, N.N.; Eitan, E.; Green, J.; Mode, N.A.; Bodogai, M.; Zhang, Y.; Lehrmann, E.; Zonderman, A.B.; Biragyn, A.; et al. Altered Extracellular Vesicle Concentration, Cargo, and Function in Diabetes. Diabetes 2018, 67, 2377–2388. [Google Scholar] [CrossRef] [Green Version]
  67. Burbidge, K.; Zwikelmaier, V.; Cook, B.; Long, M.M.; Balva, B.; Lonigro, M.; Ispas, G.; Rademacher, D.J.; Campbell, E.M. Cargo and Cell-Specific Differences in Extracellular Vesicle Populations Identified by Multiplexed Immunofluorescent Analysis. J. Extracell. Vesicles 2020, 9, 1789326. [Google Scholar] [CrossRef]
  68. Li, Y.; Zhang, H.; Du, Y.; Peng, L.; Qin, Y.; Liu, H.; Ma, X.; Wei, Y. Extracellular Vesicle MicroRNA Cargoes from Intermittent Hypoxia-Exposed Cardiomyocytes and Their Effect on Endothelium. Biochem. Bioph. Res. Commun. 2021, 548, 182–188. [Google Scholar] [CrossRef]
  69. Willms, E.; Cabañas, C.; Mäger, I.; Wood, M.J.A.; Vader, P. Extracellular Vesicle Heterogeneity: Subpopulations, Isolation Techniques, and Diverse Functions in Cancer Progression. Front. Immunol. 2018, 9, 738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Vietri, M.; Radulovic, M.; Stenmark, H. The Many Functions of ESCRTs. Nat. Rev. Mol. Cell Biol. 2020, 21, 25–42. [Google Scholar] [CrossRef]
  71. Tahir, T. Rabs Mediated Membrane Trafficking in Cancer Progression. Digit. Med. Health Technol. 2022, 2022, 1–11. [Google Scholar] [CrossRef]
  72. Mathieu, M.; Martin-Jaular, L.; Lavieu, G.; Théry, C. Specificities of Secretion and Uptake of Exosomes and Other Extracellular Vesicles for Cell-to-Cell Communication. Nat. Cell Biol. 2019, 21, 9–17. [Google Scholar] [CrossRef]
  73. Pollet, H.; Conrard, L.; Cloos, A.-S.; Tyteca, D. Plasma Membrane Lipid Domains as Platforms for Vesicle Biogenesis and Shedding? Biomolecules 2018, 8, 94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Escola, J.-M.; Kleijmeer, M.J.; Stoorvogel, W.; Griffith, J.M.; Yoshie, O.; Geuze, H.J. Selective Enrichment of Tetraspan Proteins on the Internal Vesicles of Multivesicular Endosomes and on Exosomes Secreted by Human B-Lymphocytes. J. Biol. Chem. 1998, 273, 20121–20127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Toribio, V.; Yáñez-Mó, M. Tetraspanins Interweave EV Secretion, Endosomal Network Dynamics and Cellular Metabolism. Eur. J. Cell Biol. 2022, 101, 151229. [Google Scholar] [CrossRef] [PubMed]
  76. Hemler, M.E. Tetraspanin proteins mediate cellular penetration, invasion, and fusion events and define a novel type of membrane microdomain. Annu. Rev. Cell Dev. Biol. 2003, 19, 397–422. [Google Scholar] [CrossRef] [PubMed]
  77. Perez-Hernandez, D.; Gutiérrez-Vázquez, C.; Jorge, I.; López-Martín, S.; Ursa, A.; Sánchez-Madrid, F.; Vázquez, J.; Yáñez-Mó, M. The Intracellular Interactome of Tetraspanin-Enriched Microdomains Reveals Their Function as Sorting Machineries toward Exosomes. J. Biol. Chem. 2013, 288, 11649–11661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Andreu, Z.; Yáñez-Mó, M. Tetraspanins in Extracellular Vesicle Formation and Function. Front. Immunol. 2014, 5, 442. [Google Scholar] [CrossRef] [Green Version]
  79. Gurung, S.; Perocheau, D.; Touramanidou, L.; Baruteau, J. The Exosome Journey: From Biogenesis to Uptake and Intracellular Signalling. Cell Commun. Signal. 2021, 19, 47. [Google Scholar] [CrossRef]
  80. Sahu, R.; Kaushik, S.; Clement, C.C.; Cannizzo, E.S.; Scharf, B.; Follenzi, A.; Potolicchio, I.; Nieves, E.; Cuervo, A.M.; Santambrogio, L. Microautophagy of Cytosolic Proteins by Late Endosomes. Dev. Cell 2011, 20, 131–139. [Google Scholar] [CrossRef] [Green Version]
  81. O’Brien, K.; Breyne, K.; Ughetto, S.; Laurent, L.C.; Breakefield, X.O. RNA Delivery by Extracellular Vesicles in Mammalian Cells and Its Applications. Nat. Rev. Mol. Cell Biol. 2020, 21, 585–606. [Google Scholar] [CrossRef]
  82. Turchinovich, A.; Drapkina, O.; Tonevitsky, A. Transcriptome of Extracellular Vesicles: State-of-the-Art. Front. Immunol. 2019, 10, 202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Thakur, B.K.; Zhang, H.; Becker, A.; Matei, I.; Huang, Y.; Costa-Silva, B.; Zheng, Y.; Hoshino, A.; Brazier, H.; Xiang, J.; et al. Double-Stranded DNA in Exosomes: A Novel Biomarker in Cancer Detection. Cell Res. 2014, 24, 766–769. [Google Scholar] [CrossRef] [Green Version]
  84. Sansone, P.; Savini, C.; Kurelac, I.; Chang, Q.; Amato, L.B.; Strillacci, A.; Stepanova, A.; Iommarini, L.; Mastroleo, C.; Daly, L.; et al. Packaging and Transfer of Mitochondrial DNA via Exosomes Regulate Escape from Dormancy in Hormonal Therapy-Resistant Breast Cancer. Proc. Natl. Acad. Sci. USA 2017, 114, E9066–E9075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Vagner, T.; Spinelli, C.; Minciacchi, V.R.; Balaj, L.; Zandian, M.; Conley, A.; Zijlstra, A.; Freeman, M.R.; Demichelis, F.; De, S.; et al. Large Extracellular Vesicles Carry Most of the Tumour DNA Circulating in Prostate Cancer Patient Plasma. J. Extracell. Vesicles 2018, 7, 1505403. [Google Scholar] [CrossRef] [Green Version]
  86. Shelke, G.; Jang, S.C.; Yin, Y.; Lässer, C.; Lötvall, J. Human Mast Cells Release Extracellular Vesicle-Associated DNA. Matters 2016, 2, e201602000034. [Google Scholar] [CrossRef] [Green Version]
  87. Fernando, M.R.; Jiang, C.; Krzyzanowski, G.D.; Ryan, W.L. New Evidence That a Large Proportion of Human Blood Plasma Cell-Free DNA Is Localized in Exosomes. PLoS ONE 2017, 12, e0183915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Kujala, J.; Hartikainen, J.M.; Tengström, M.; Sironen, R.; Kosma, V.; Mannermaa, A. High Mutation Burden of Circulating Cell-free DNA in Early-stage Breast Cancer Patients Is Associated with a Poor Relapse-free Survival. Cancer Med. 2020, 9, 5922–5931. [Google Scholar] [CrossRef]
  89. Alimirzaie, S.; Bagherzadeh, M.; Akbari, M.R. Liquid Biopsy in Breast Cancer: A Comprehensive Review. Clin. Genet. 2019, 95, 643–660. [Google Scholar] [CrossRef]
  90. Hall, C.; Sarli, V.; Meas, S.; Lucci, A. Role of Liquid Biopsy in Clinical Decision-Making for Breast Cancer. Curr. Breast Cancer Rep. 2019, 11, 52–66. [Google Scholar] [CrossRef]
  91. Zebrowska, A.; Skowronek, A.; Wojakowska, A.; Widlak, P.; Pietrowska, M. Metabolome of Exosomes: Focus on Vesicles Released by Cancer Cells and Present in Human Body Fluids. Int. J. Mol. Sci. 2019, 20, 3461. [Google Scholar] [CrossRef] [Green Version]
  92. Skotland, T.; Sandvig, K.; Llorente, A. Lipids in Exosomes: Current Knowledge and the Way Forward. Prog. Lipid Res. 2017, 66, 30–41. [Google Scholar] [CrossRef]
  93. Fonseca, P.; Vardaki, I.; Occhionero, A.; Panaretakis, T. Chapter Five Metabolic and Signaling Functions of Cancer Cell-Derived Extracellular Vesicles. Int. Rev. Cell Mol. Biol. 2016, 326, 175–199. [Google Scholar] [CrossRef] [PubMed]
  94. Tkach, M.; Kowal, J.; Théry, C. Why the Need and How to Approach the Functional Diversity of Extracellular Vesicles. Phil-Osophical Trans. R. Soc. B Biol. Sci. 2018, 373, 20160479. [Google Scholar] [CrossRef] [PubMed]
  95. Maisano, D.; Mimmi, S.; Russo, R.; Fioravanti, A.; Fiume, G.; Vecchio, E.; Nisticò, N.; Quinto, I.; Iaccino, E. Uncovering the Exosomes Diversity: A Window of Opportunity for Tumor Progression Monitoring. Pharmaceuticals 2020, 13, 180. [Google Scholar] [CrossRef]
  96. Zabeo, D.; Cvjetkovic, A.; Lässer, C.; Schorb, M.; Lötvall, J.; Höög, J.L. Exosomes Purified from a Single Cell Type Have Di-verse Morphology. J. Extracell. Vesicles 2017, 6, 1329476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Vagner, T.; Chin, A.; Mariscal, J.; Bannykh, S.; Engman, D.M.; Vizio, D.D. Protein Composition Reflects Extracellular Vesicle Heterogeneity. Proteomics 2019, 19, e1800167. [Google Scholar] [CrossRef]
  98. Montecchi, T.; Shaba, E.; Tommaso, D.D.; Giuseppe, F.D.; Angelucci, S.; Bini, L.; Landi, C.; Baldari, C.T.; Ulivieri, C. Differen-tial Proteomic Analysis of Astrocytes and Astrocytes-Derived Extracellular Vesicles from Control and Rai Knockout Mice: In-sights into the Mechanisms of Neuroprotection. Int. J. Mol. Sci. 2021, 22, 7933. [Google Scholar] [CrossRef]
  99. Sork, H.; Corso, G.; Krjutskov, K.; Johansson, H.J.; Nordin, J.Z.; Wiklander, O.P.B.; Lee, Y.X.F.; Westholm, J.O.; Lehtiö, J.; Wood, M.J.A.; et al. Heterogeneity and Interplay of the Extracellular Vesicle Small RNA Transcriptome and Proteome. Sci. Rep. 2018, 8, 10813. [Google Scholar] [CrossRef] [Green Version]
  100. Shaba, E.; Vantaggiato, L.; Governini, L.; Haxhiu, A.; Sebastiani, G.; Fignani, D.; Grieco, G.E.; Bergantini, L.; Bini, L.; Landi, C. Multi-Omics Integrative Approach of Extracellular Vesicles: A Future Challenging Milestone. Proteomes 2022, 10, 12. [Google Scholar] [CrossRef]
  101. Greening, D.W.; Simpson, R.J. Understanding Extracellular Vesicle Diversity—Current Status. Expert Rev. Proteomic 2018, 15, 887–910. [Google Scholar] [CrossRef] [PubMed]
  102. Kolonics, F.; Szeifert, V.; Timár, C.I.; Ligeti, E.; Lőrincz, Á.M. The Functional Heterogeneity of Neutrophil-Derived Extra-cellular Vesicles Reflects the Status of the Parent Cell. Cells 2020, 9, 2718. [Google Scholar] [CrossRef] [PubMed]
  103. Carnino, J.M.; Ni, K.; Jin, Y. Post-Translational Modification Regulates Formation and Cargo-Loading of Extracellular Vesicles. Front. Immunol. 2020, 11, 948. [Google Scholar] [CrossRef] [PubMed]
  104. Wei, H.; Chen, Q.; Lin, L.; Sha, C.; Li, T.; Liu, Y.; Yin, X.; Xu, Y.; Chen, L.; Gao, W.; et al. Regulation of Exosome Production and Cargo Sorting. Int. J. Biol. Sci. 2021, 17, 163–177. [Google Scholar] [CrossRef]
  105. Riches, A.; Campbell, E.; Borger, E.; Powis, S. Regulation of Exosome Release from Mammary Epithelial and Breast Cancer Cells—A New Regulatory Pathway. Eur. J. Cancer 2014, 50, 1025–1034. [Google Scholar] [CrossRef] [PubMed]
  106. Muntasell, A.; Berger, A.C.; Roche, P.A. T cell-induced secretion of MHC class II–peptide complexes on B cell exosomes. EMBO J. 2007, 26, 4263–4272. [Google Scholar] [CrossRef] [Green Version]
  107. Takahashi, Y.; Nishikawa, M.; Shinotsuka, H.; Matsui, Y.; Ohara, S.; Imai, T.; Takakura, Y. Visualization and in Vivo Tracking of the Exosomes of Murine Melanoma B16-BL6 Cells in Mice after Intravenous Injection. J. Biotechnol. 2013, 165, 77–84. [Google Scholar] [CrossRef]
  108. Zech, D.; Rana, S.; Büchler, M.W.; Zöller, M. Tumor-Exosomes and Leukocyte Activation: An Ambivalent Crosstalk. Cell Commun. Signal. 2012, 10, 37. [Google Scholar] [CrossRef] [Green Version]
  109. Horibe, S.; Tanahashi, T.; Kawauchi, S.; Murakami, Y.; Rikitake, Y. Mechanism of Recipient Cell-Dependent Differences in Exosome Uptake. BMC Cancer 2018, 18, 47. [Google Scholar] [CrossRef] [Green Version]
  110. Joshi, B.S.; de Beer, M.A.; Giepmans, B.N.G.; Zuhorn, I.S. Endocytosis of Extracellular Vesicles and Release of Their Cargo from Endosomes. ACS Nano 2020, 14, 4444–4455. [Google Scholar] [CrossRef] [Green Version]
  111. Gonda, A.; Kabagwira, J.; Senthil, G.N.; Wall, N.R. Internalization of Exosomes through Receptor-Mediated Endocytosis. Mol. Cancer Res. 2018, 17, 337–347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Kwok, Z.H.; Wang, C.; Jin, Y. Extracellular Vesicle Transportation and Uptake by Recipient Cells: A Critical Process to Regulate Human Diseases. Process 2021, 9, 273. [Google Scholar] [CrossRef] [PubMed]
  113. Dzobo, K.; Senthebane, D.A.; Ganz, C.; Thomford, N.E.; Wonkam, A.; Dandara, C. Advances in Therapeutic Targeting of Cancer Stem Cells within the Tumor Microenvironment: An Updated Review. Cells 2020, 9, 1896. [Google Scholar] [CrossRef]
  114. Zhang, X.; Liu, D.; Gao, Y.; Lin, C.; An, Q.; Feng, Y.; Liu, Y.; Liu, D.; Luo, H.; Wang, D. The Biology and Function of Extracellular Vesicles in Cancer Development. Front. Cell Dev. Biol. 2021, 9, 777441. [Google Scholar] [CrossRef]
  115. Bissell, M.J.; Rizki, A.; Mian, I.S. Tissue Architecture: The Ultimate Regulator of Breast Epithelial Function. Curr. Opin. Cell Biol. 2003, 15, 753–762. [Google Scholar] [CrossRef] [Green Version]
  116. Daniel, C.W.; Smith, G.H. The Mammary Gland: A Model for Development. J. Mammary Gland. Biol. 1999, 4, 3–8. [Google Scholar] [CrossRef]
  117. Krause, S.; Maffini, M.V.; Soto, A.M.; Sonnenschein, C. A Novel 3D In Vitro Culture Model to Study StromalEpithelial Interactions in the Mammary Gland. Tissue Eng. Part C Methods 2008, 14, 261–271. [Google Scholar] [CrossRef] [Green Version]
  118. Howard, B.A.; Lu, P. Stromal Regulation of Embryonic and Postnatal Mammary Epithelial Development and Differentiation. Semin. Cell Dev. Biol. 2014, 25, 43–51. [Google Scholar] [CrossRef] [PubMed]
  119. Schedin, P.; Hovey, R.C. Editorial: The Mammary Stroma in Normal Development and Function. J. Mammary Gland. Biol. 2010, 15, 275–277. [Google Scholar] [CrossRef] [Green Version]
  120. Ronnov-Jessen, L.; Petersen, O.W.; Bissell, M.J. Cellular Changes Involved in Conversion of Normal to Malignant Breast: Importance of the Stromal Reaction. Physiol. Rev. 1996, 76, 69–125. [Google Scholar] [CrossRef] [PubMed]
  121. Welsch, U.; Oppermann, T.; Mortezza, M.; Höfter, E.; Unterberger, P. Secretory Phenomena in the Non-Lactating Human Mammary Gland. Ann. Anat.-Anat. Anz. 2007, 189, 131–141. [Google Scholar] [CrossRef] [PubMed]
  122. Nakatani, H.; Aoki, N.; Nakagawa, Y.; Jin-No, S.; Aoyama, K.; Oshima, K.; Ohira, S.; Sato, C.; Nadano, D.; Matsuda, T. Weaning-Induced Expression of a Milk-Fat Globule Protein, MFG-E8, in Mouse Mammary Glands, as Demonstrated by the Analyses of Its MRNA, Protein and Phosphatidylserine-Binding Activity. Biochem. J. 2006, 395, 21–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Conklin, M.W.; Keely, P.J. Why the Stroma Matters in Breast Cancer: Insights into Breast Cancer Patient Outcomes through the Examination of Stromal Biomarkers. Cell Adhes. Migr. 2012, 6, 249–260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Pujuguet, P.; Simian, M.; Liaw, J.; Timpl, R.; Werb, Z.; Bissell, M.J. Nidogen-1 Regulates Laminin-1-Dependent Mammary-Specific Gene Expression. J. Cell Sci. 2000, 113, 849–858. [Google Scholar] [CrossRef] [PubMed]
  125. Hendrix, A.; Hume, A.N. Exosome Signaling in Mammary Gland Development and Cancer. Int. J. Dev. Biol. 2011, 55, 879–887. [Google Scholar] [CrossRef] [Green Version]
  126. Lin, M.-C.; Chen, S.-Y.; He, P.-L.; Luo, W.-T.; Li, H.-J. Transfer of Mammary Gland-Forming Ability Between Mammary Basal Epithelial Cells and Mammary Luminal Cells via Extracellular Vesicles/Exosomes. J. Vis. Exp. 2017, 124, e55736. [Google Scholar] [CrossRef]
  127. Lakkaraju, A.; Rodriguez-Boulan, E. Itinerant exosomes: Emerging roles in cell and tissue polarity. Trends Cell Biol. 2008, 18, 199–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Chin, A.R.; Yan, W.; Cao, M.; Liu, X.; Wang, S.E. Polarized Secretion of Extracellular Vesicles by Mammary Epithelia. J. Mammary Gland. Biol. 2018, 23, 165–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Chen, Q.; Takada, R.; Noda, C.; Kobayashi, S.; Takada, S. Different Populations of Wnt-Containing Vesicles Are Individually Released from Polarized Epithelial Cells. Sci. Rep. 2016, 6, 35562. [Google Scholar] [CrossRef] [Green Version]
  130. Colombo, F.; Casella, G.; Podini, P.; Finardi, A.; Racchetti, G.; Norton, E.G.; Cocucci, E.; Furlan, R. Polarized Cells Display Asymmetric Release of Extracellular Vesicles. Traffic 2021, 22, 98–110. [Google Scholar] [CrossRef]
  131. Sreekumar, P.G.; Kannan, R.; Kitamura, M.; Spee, C.; Barron, E.; Ryan, S.J.; Hinton, D.R. αB Crystallin Is Apically Secreted within Exosomes by Polarized Human Retinal Pigment Epithelium and Provides Neuroprotection to Adjacent Cells. PLoS ONE 2010, 5, e12578. [Google Scholar] [CrossRef] [Green Version]
  132. Muschler, J.; Streuli, C.H. Cell–Matrix Interactions in Mammary Gland Development and Breast Cancer. Cold Spring Harb. Perspect. Biol. 2010, 2, a003202. [Google Scholar] [CrossRef]
  133. Sternlicht, M.D.; Kouros-Mehr, H.; Lu, P.; Werb, Z. Hormonal and Local Control of Mammary Branching Morphogenesis. Differentiation 2006, 74, 365–381. [Google Scholar] [CrossRef] [Green Version]
  134. Fu, N.Y.; Nolan, E.; Lindeman, G.J.; Visvader, J.E. Stem Cells and the Differentiation Hierarchy in Mammary Gland Development. Physiol. Rev. 2020, 100, 489–523. [Google Scholar] [CrossRef]
  135. Fu, N.Y.; Rios, A.C.; Pal, B.; Law, C.W.; Jamieson, P.; Liu, R.; Vaillant, F.; Jackling, F.; Liu, K.H.; Smyth, G.K.; et al. Identification of Quiescent and Spatially Restricted Mammary Stem Cells That Are Hormone Responsive. Nat. Cell Biol. 2017, 19, 164–176. [Google Scholar] [CrossRef] [PubMed]
  136. Stingl, J.; Eirew, P.; Ricketson, I.; Shackleton, M.; Vaillant, F.; Choi, D.; Li, H.I.; Eaves, C.J. Purification and Unique Properties of Mammary Epithelial Stem Cells. Nature 2006, 439, 993–997. [Google Scholar] [CrossRef] [PubMed]
  137. Cai, S.; Kalisky, T.; Sahoo, D.; Dalerba, P.; Feng, W.; Lin, Y.; Qian, D.; Kong, A.; Yu, J.; Wang, F.; et al. A Quiescent Bcl11b High Stem Cell Population Is Required for Maintenance of the Mammary Gland. Cell Stem Cell 2017, 20, 247–260.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Zeng, Y.A.; Nusse, R. Wnt Proteins Are Self-Renewal Factors for Mammary Stem Cells and Promote Their Long-Term Expansion in Culture. Cell Stem Cell 2010, 6, 568–577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Zhao, C.; Cai, S.; Shin, K.; Lim, A.; Kalisky, T.; Lu, W.-J.; Clarke, M.F.; Beachy, P.A. Stromal Gli2 Activity Coordinates a Niche Signaling Program for Mammary Epithelial Stem Cells. Science 2017, 356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Pardal, R.; Clarke, M.F.; Morrison, S.J. Applying the Principles of Stem-Cell Biology to Cancer. Nat. Rev. Cancer 2003, 3, 895–902. [Google Scholar] [CrossRef]
  141. Abdelhamed, S.; Butler, J.T.; Doron, B.; Halse, A.; Nemecek, E.; Wilmarth, P.A.; Marks, D.L.; Chang, B.H.; Horton, T.; Kurre, P. Extracellular Vesicles Impose Quiescence on Residual Hematopoietic Stem Cells in the Leukemic Niche. EMBO Rep. 2019, 20, e47546. [Google Scholar] [CrossRef]
  142. Victora, C.G.; Bahl, R.; Barros, A.J.D.; França, G.V.A.; Horton, S.; Krasevec, J.; Murch, S.; Sankar, M.J.; Walker, N.; Rollins, N.C.; et al. Breastfeeding in the 21st Century: Epidemiology, Mechanisms, and Lifelong Effect. Lancet 2016, 387, 475–490. [Google Scholar] [CrossRef] [Green Version]
  143. Liao, Y.; Du, X.; Li, J.; Lönnerdal, B. Human Milk Exosomes and Their MicroRNAs Survive Digestion in Vitro and Are Taken up by Human Intestinal Cells. Mol. Nutr. Food Res. 2017, 61, 1700082. [Google Scholar] [CrossRef] [PubMed]
  144. Kusuma, R.J.; Manca, S.; Friemel, T.; Sukreet, S.; Nguyen, C.; Zempleni, J. Human Vascular Endothelial Cells Transport Foreign Exosomes from Cow’s Milk by Endocytosis. Am. J. Physiol.-Cell Physiol. 2016, 310, C800–C807. [Google Scholar] [CrossRef] [Green Version]
  145. Jiang, X.; You, L.; Zhang, Z.; Cui, X.; Zhong, H.; Sun, X.; Ji, C.; Chi, X. Biological Properties of Milk-Derived Extracellular Vesicles and Their Physiological Functions in Infant. Front. Cell Dev. Biol. 2021, 9, 693534. [Google Scholar] [CrossRef] [PubMed]
  146. Paredes, P.T.; Gutzeit, C.; Johansson, S.; Admyre, C.; Stenius, F.; Alm, J.; Scheynius, A.; Gabrielsson, S. Differences in Exosome Populations in Human Breast Milk in Relation to Allergic Sensitization and Lifestyle. Allergy 2014, 69, 463–471. [Google Scholar] [CrossRef] [PubMed]
  147. Zonneveld, M.I.; van Herwijnen, M.J.C.; Fernandez-Gutierrez, M.M.; Giovanazzi, A.; de Groot, A.M.; Kleinjan, M.; van Capel, T.M.M.; Sijts, A.J.A.M.; Taams, L.S.; Garssen, J.; et al. Human Milk Extracellular Vesicles Target Nodes in Interconnected Signalling Pathways That Enhance Oral Epithelial Barrier Function and Dampen Immune Responses. J. Extracell. Vesicles 2021, 10, e12071. [Google Scholar] [CrossRef] [PubMed]
  148. Zeng, B.; Chen, T.; Luo, J.-Y.; Zhang, L.; Xi, Q.-Y.; Jiang, Q.-Y.; Sun, J.-J.; Zhang, Y.-L. Biological Characteristics and Roles of Noncoding RNAs in Milk-Derived Extracellular Vesicles. Adv. Nutr. 2021, 12, 1006–1019. [Google Scholar] [CrossRef]
  149. Aguilar-Lozano, A.; Baier, S.; Grove, R.; Shu, J.; Giraud, D.; Leiferman, A.; Mercer, K.E.; Cui, J.; Badger, T.M.; Adamec, J.; et al. Concentrations of Purine Metabolites Are Elevated in Fluids from Adults and Infants and in Livers from Mice Fed Diets Depleted of Bovine Milk Exosomes and Their RNA Cargos. J. Nutr. 2018, 148, 1886–1894. [Google Scholar] [CrossRef] [Green Version]
  150. Nichols, H.B.; Schoemaker, M.J.; Cai, J.; Xu, J.; Wright, L.B.; Brook, M.N.; Jones, M.E.; Adami, H.-O.; Baglietto, L.; Bertrand, K.A.; et al. Breast Cancer Risk After Recent Childbirth: A Pooled Analysis of 15 Prospective Studies. Ann. Intern. Med. 2018, 170, 22. [Google Scholar] [CrossRef] [Green Version]
  151. Sauter, E.R.; Reidy, D. How Exosomes in Human Breast Milk May Influence Breast Cancer Risk. Transl. Cancer Res. 2017, 6, S1384–S1388. [Google Scholar] [CrossRef]
  152. Qin, W.; Tsukasaki, Y.; Dasgupta, S.; Mukhopadhyay, N.; Ikebe, M.; Sauter, E.R. Exosomes in Human Breast Milk Promote EMT. Clin. Cancer Res. 2016, 22, 4517–4524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Mallini, P.; Lennard, T.; Kirby, J.; Meeson, A. Epithelial-to-Mesenchymal Transition: What Is the Impact on Breast Cancer Stem Cells and Drug Resistance. Cancer Treat. Rev. 2014, 40, 341–348. [Google Scholar] [CrossRef] [PubMed]
  154. Ramamoorthi, G.; Kodumudi, K.; Gallen, C.; Zachariah, N.N.; Basu, A.; Albert, G.; Beyer, A.; Snyder, C.; Wiener, D.; Costa, R.L.B.; et al. Disseminated Cancer Cells in Breast Cancer: Mechanism of Dissemination and Dormancy and Emerging Insights on Therapeutic Opportunities. Semin. Cancer Biol. 2022, 78, 78–89. [Google Scholar] [CrossRef]
  155. Harmes, D.C.; DiRenzo, J. Cellular Quiescence in Mammary Stem Cells and Breast Tumor Stem Cells: Got Testable Hypotheses? J. Mammary Gland. Biol. 2009, 14, 19–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Lin, H. Cell Biology of Stem Cells: An Enigma of Asymmetry and Self-Renewal. J. Cell Biol. 2008, 180, 257–260. [Google Scholar] [CrossRef] [Green Version]
  157. Guen, V.J.; Chavarria, T.E.; Kröger, C.; Ye, X.; Weinberg, R.A.; Lees, J.A. EMT Programs Promote Basal Mammary Stem Cell and Tumor-Initiating Cell Stemness by Inducing Primary Ciliogenesis and Hedgehog Signaling. Proc. Natl. Acad. Sci. USA 2017, 114, E10532–E10539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Zhu, R.; Gires, O.; Zhu, L.; Liu, J.; Li, J.; Yang, H.; Ju, G.; Huang, J.; Ge, W.; Chen, Y.; et al. TSPAN8 Promotes Cancer Cell Stemness via Activation of Sonic Hedgehog Signaling. Nat. Commun. 2019, 10, 2863. [Google Scholar] [CrossRef] [Green Version]
  159. Giuli, M.V.; Giuliani, E.; Screpanti, I.; Bellavia, D.; Checquolo, S. Notch Signaling Activation as a Hallmark for Triple-Negative Breast Cancer Subtype. J. Oncol. 2019, 2019, 8707053. [Google Scholar] [CrossRef]
  160. Al-Hajj, M.; Wicha, M.S.; Benito-Hernandez, A.; Morrison, S.J.; Clarke, M.F. Prospective Identification of Tumorigenic Breast Cancer Cells. Proc. Natl. Acad. Sci. USA 2003, 100, 3983–3988. [Google Scholar] [CrossRef] [Green Version]
  161. Woodward, W.A.; Chen, M.S.; Behbod, F.; Rosen, J.M. On Mammary Stem Cells. J. Cell Sci. 2005, 118, 3585–3594. [Google Scholar] [CrossRef] [Green Version]
  162. Vuong, D.; Simpson, P.T.; Green, B.; Cummings, M.C.; Lakhani, S.R. Molecular Classification of Breast Cancer. Virchows Arch. 2014, 465, 1–14. [Google Scholar] [CrossRef]
  163. Parmar, H.; Cunha, G.R. Epithelial–Stromal Interactions in the Mouse and Human Mammary Gland in Vivo. Endocr.-Relat. Cancer 2004, 11, 437–458. [Google Scholar] [CrossRef] [Green Version]
  164. Ingthorsson, S.; Briem, E.; Bergthorsson, J.T.; Gudjonsson, T. Epithelial Plasticity During Human Breast Morphogenesis and Cancer Progression. J. Mammary Gland. Biol. 2016, 21, 139–148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Sansone, P.; Ceccarelli, C.; Berishaj, M.; Chang, Q.; Rajasekhar, V.K.; Perna, F.; Bowman, R.L.; Vidone, M.; Daly, L.; Nnoli, J.; et al. Self-Renewal of CD133(Hi) Cells by IL6/Notch3 Signalling Regulates Endocrine Resistance in Metastatic Breast Cancer. Nat. Commun. 2016, 7, 10442. [Google Scholar] [CrossRef] [Green Version]
  166. Ratajczak, J.; Miekus, K.; Kucia, M.; Zhang, J.; Reca, R.; Dvorak, P.; Ratajczak, M.Z. Embryonic stem cell-derived microvesicles reprogram hematopoietic progenitors: Evidence for horizontal transfer of mRNA and protein delivery. Leukemia 2006, 20, 847–856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Badve, S.; Nakshatri, H. Breast-Cancer Stem Cells—Beyond Semantics. Lancet Oncol. 2012, 13, e43–e48. [Google Scholar] [CrossRef] [PubMed]
  168. Lin, M.; Chen, S.; Tsai, H.; He, P.; Lin, Y.; Herschman, H.; Li, H. PGE2/EP4 Signaling Controls the Transfer of the Mammary Stem Cell State by Lipid Rafts in Extracellular Vesicles. Stem Cells 2017, 35, 425–444. [Google Scholar] [CrossRef]
  169. Kong, X.; Zhang, J.; Li, J.; Shao, J.; Fang, L. MiR-130a-3p Inhibits Migration and Invasion by Regulating RAB5B in Human Breast Cancer Stem Cell-like Cells. Biochem. Bioph. Res. Commun. 2018, 501, 486–493. [Google Scholar] [CrossRef]
  170. Visvader, J.E.; Lindeman, G.J. Cancer Stem Cells in Solid Tumours: Accumulating Evidence and Unresolved Questions. Nat. Rev. Cancer 2008, 8, 755–768. [Google Scholar] [CrossRef]
  171. Bliss, S.A.; Sinha, G.; Sandiford, O.A.; Williams, L.M.; Engelberth, D.J.; Guiro, K.; Isenalumhe, L.L.; Greco, S.J.; Ayer, S.; Bryan, M.; et al. Mesenchymal Stem Cell–Derived Exosomes Stimulate Cycling Quiescence and Early Breast Cancer Dormancy in Bone Marrow. Cancer Res. 2016, 76, 5832–5844. [Google Scholar] [CrossRef] [Green Version]
  172. Shen, M.; Dong, C.; Ruan, X.; Yan, W.; Cao, M.; Pizzo, D.; Wu, X.; Yang, L.; Liu, L.; Ren, X.; et al. Chemotherapy-Induced Extracellular Vesicle MiRNAs Promote Breast Cancer Stemness by Targeting ONECUT2. Cancer Res. 2019, 79, 3608–3621. [Google Scholar] [CrossRef]
  173. Ciardiello, C.; Leone, A.; Budillon, A. The Crosstalk between Cancer Stem Cells and Microenvironment Is Critical for Solid Tumor Progression: The Significant Contribution of Extracellular Vesicles. Stem Cells Int. 2018, 2018, 6392198. [Google Scholar] [CrossRef] [PubMed]
  174. Chou, J.; Shahi, P.; Werb, Z. MicroRNA-Mediated Regulation of the Tumor Microenvironment. Cell Cycle 2013, 12, 3262–3271. [Google Scholar] [CrossRef] [Green Version]
  175. Leal-Orta, E.; Ramirez-Ricardo, J.; Cortes-Reynosa, P.; Galindo-Hernandez, O.; Salazar, E.P. Role of PI3K/Akt on Migration and Invasion of MCF10A Cells Treated with Extracellular Vesicles from MDA-MB-231 Cells Stimulated with Linoleic Acid. J. Cell Commun. Signal. 2019, 13, 235–244. [Google Scholar] [CrossRef]
  176. Ozawa, P.M.M.; Alkhilaiwi, F.; Cavalli, I.J.; Malheiros, D.; Ribeiro, E.M.d.S.F.; Cavalli, L.R. Extracellular Vesicles from Triple-Negative Breast Cancer Cells Promote Proliferation and Drug Resistance in Non-Tumorigenic Breast Cells. Breast Cancer Res. Treat. 2018, 172, 713–723. [Google Scholar] [CrossRef]
  177. Ren, Z.; Lv, M.; Yu, Q.; Bao, J.; Lou, K.; Li, X. MicroRNA-370-3p Shuttled by Breast Cancer Cell-derived Extracellular Vesicles Induces Fibroblast Activation through the CYLD/Nf-κB Axis to Promote Breast Cancer Progression. FASEB J. 2021, 35, e21383. [Google Scholar] [CrossRef]
  178. Naito, Y.; Yamamoto, Y.; Sakamoto, N.; Shimomura, I.; Kogure, A.; Kumazaki, M.; Yokoi, A.; Yashiro, M.; Kiyono, T.; Yanagihara, K.; et al. Cancer Extracellular Vesicles Contribute to Stromal Heterogeneity by Inducing Chemokines in Cancer-Associated Fibroblasts. Oncogene 2019, 38, 5566–5579. [Google Scholar] [CrossRef]
  179. Baroni, S.; Romero-Cordoba, S.; Plantamura, I.; Dugo, M.; D’Ippolito, E.; Cataldo, A.; Cosentino, G.; Angeloni, V.; Rossini, A.; Daidone, M.G.; et al. Exosome-Mediated Delivery of MiR-9 Induces Cancer-Associated Fibroblast-like Properties in Human Breast Fibroblasts. Cell Death Dis. 2016, 7, e2312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Zhou, W.; Fong, M.Y.; Min, Y.; Somlo, G.; Liu, L.; Palomares, M.R.; Yu, Y.; Chow, A.; O’Connor, S.T.F.; Chin, A.R.; et al. Cancer-Secreted MiR-105 Destroys Vascular Endothelial Barriers to Promote Metastasis. Cancer Cell 2014, 25, 501–515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Rabe, D.C.; Rustandy, F.D.; Lee, J.; Rosner, M.R. Tumor Extracellular Vesicles Are Required for Tumor-Associated Macrophage Programming. Biorxiv 2018, 375022. [Google Scholar] [CrossRef] [Green Version]
  182. Tkach, M.; Thalmensi, J.; Timperi, E.; Gueguen, P.; Névo, N.; Grisard, E.; Sirven, P.; Cocozza, F.; Gouronnec, A.; Martin-Jaular, L.; et al. Extracellular Vesicles from Triple Negative Breast Cancer Promote Pro-Inflammatory Macrophages Associated with Better Clinical Outcome. Proc. Natl. Acad. Sci. USA 2022, 119, e2107394119. [Google Scholar] [CrossRef]
  183. Xun, J.; Du, L.; Gao, R.; Shen, L.; Wang, D.; Kang, L.; Chen, C.; Zhang, Z.; Zhang, Y.; Yue, S.; et al. Cancer-Derived Exosomal MiR-138-5p Modulates Polarization of Tumor-Associated Macrophages through Inhibition of KDM6B. Theranostics 2021, 11, 6847–6859. [Google Scholar] [CrossRef] [PubMed]
  184. Liang, Y.; Song, X.; Li, Y.; Chen, B.; Zhao, W.; Wang, L.; Zhang, H.; Liu, Y.; Han, D.; Zhang, N.; et al. LncRNA BCRT1 Promotes Breast Cancer Progression by Targeting MiR-1303/PTBP3 Axis. Mol. Cancer 2020, 19, 85. [Google Scholar] [CrossRef]
  185. Ham, S.; Lima, L.G.; Chai, E.P.Z.; Muller, A.; Lobb, R.J.; Krumeich, S.; Wen, S.W.; Wiegmans, A.P.; Möller, A. Breast Cancer-Derived Exosomes Alter Macrophage Polarization via Gp130/STAT3 Signaling. Front. Immunol. 2018, 9, 871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Maji, S.; Chaudhary, P.; Akopova, I.; Nguyen, P.M.; Hare, R.J.; Gryczynski, I.; Vishwanatha, J.K. Exosomal Annexin II Promotes Angiogenesis and Breast Cancer Metastasis. Mol. Cancer Res. 2017, 15, 93–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Dumas, J.-F.; Brisson, L. Interaction between Adipose Tissue and Cancer Cells: Role for Cancer Progression. Cancer Metast. Rev. 2021, 40, 31–46. [Google Scholar] [CrossRef]
  188. Lapeire, L.; Hendrix, A.; Lambein, K.; Bockstal, M.V.; Braems, G.; Broecke, R.V.D.; Limame, R.; Mestdagh, P.; Vandesompele, J.; Vanhove, C.; et al. Cancer-Associated Adipose Tissue Promotes Breast Cancer Progression by Paracrine Oncostatin M and Jak/STAT3 Signaling. Cancer Res. 2014, 74, 6806–6819. [Google Scholar] [CrossRef] [Green Version]
  189. Wu, Q.; Li, J.; Li, Z.; Sun, S.; Zhu, S.; Wang, L.; Wu, J.; Yuan, J.; Zhang, Y.; Sun, S.; et al. Exosomes from the Tumour-Adipocyte Interplay Stimulate Beige/Brown Differentiation and Reprogram Metabolism in Stromal Adipocytes to Promote Tumour Progression. J. Exp. Clin. Cancer Res. 2019, 38, 223. [Google Scholar] [CrossRef] [Green Version]
  190. Cho, J.A.; Park, H.; Lim, E.H.; Lee, K.W. Exosomes from Breast Cancer Cells Can Convert Adipose Tissue-Derived Mesenchymal Stem Cells into Myofibroblast-like Cells. Int. J. Oncol. 2011, 40, 130–138. [Google Scholar] [CrossRef] [Green Version]
  191. Bao, Q.; Huang, Q.; Chen, Y.; Wang, Q.; Sang, R.; Wang, L.; Xie, Y.; Chen, W. Tumor-Derived Extracellular Vesicles Regulate Cancer Progression in the Tumor Microenvironment. Front. Mol. Biosci. 2022, 8, 796385. [Google Scholar] [CrossRef]
  192. Li, C.; Teixeira, A.F.; Zhu, H.-J.; Dijke, P. ten Cancer Associated-Fibroblast-Derived Exosomes in Cancer Progression. Mol. Cancer 2021, 20, 154. [Google Scholar] [CrossRef] [PubMed]
  193. Wang, H.; Wei, H.; Wang, J.; Li, L.; Chen, A.; Li, Z. MicroRNA-181d-5p-Containing Exosomes Derived from CAFs Promote EMT by Regulating CDX2/HOXA5 in Breast Cancer. Mol. Ther.-Nucleic Acids 2020, 19, 654–667. [Google Scholar] [CrossRef]
  194. Donnarumma, E.; Fiore, D.; Nappa, M.; Roscigno, G.; Adamo, A.; Iaboni, M.; Russo, V.; Affinito, A.; Puoti, I.; Quintavalle, C.; et al. Cancer-Associated Fibroblasts Release Exosomal MicroRNAs That Dictate an Aggressive Phenotype in Breast Cancer. Oncotarget 2017, 8, 19592–19608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Chen, Y.; Zeng, C.; Zhan, Y.; Wang, H.; Jiang, X.; Li, W. Aberrant Low Expression of P85α in Stromal Fibroblasts Promotes Breast Cancer Cell Metastasis through Exosome-Mediated Paracrine Wnt10b. Oncogene 2017, 36, 4692–4705. [Google Scholar] [CrossRef] [Green Version]
  196. Li, Y.; Zhao, Z.; Liu, W.; Li, X. SNHG3 Functions as MiRNA Sponge to Promote Breast Cancer Cells Growth Through the Metabolic Reprogramming. Appl. Biochem. Biotech. 2020, 191, 1084–1099. [Google Scholar] [CrossRef] [Green Version]
  197. Yang, S.-S.; Ma, S.; Dou, H.; Liu, F.; Zhang, S.-Y.; Jiang, C.; Xiao, M.; Huang, Y.-X. Breast Cancer-Derived Exosomes Regulate Cell Invasion and Metastasis in Breast Cancer via MiR-146a to Activate Cancer Associated Fibroblasts in Tumor Microenvironment. Exp. Cell Res. 2020, 391, 111983. [Google Scholar] [CrossRef]
  198. Rezaei, M.; Cavaco, A.C.M.; Stehling, M.; Nottebaum, A.; Brockhaus, K.; Caliandro, M.F.; Schelhaas, S.; Schmalbein, F.; Vestweber, D.; Eble, J.A. Extracellular Vesicle Transfer from Endothelial Cells Drives VE-Cadherin Expression in Breast Cancer Cells, Thereby Causing Heterotypic Cell Contacts. Cancers 2020, 12, 2138. [Google Scholar] [CrossRef] [PubMed]
  199. Lombardo, G.; Gili, M.; Grange, C.; Cavallari, C.; Dentelli, P.; Togliatto, G.; Taverna, D.; Camussi, G.; Brizzi, M.F. IL-3R-Alpha Blockade Inhibits Tumor Endothelial Cell-Derived Extracellular Vesicle (EV)-Mediated Vessel Formation by Targeting the β-Catenin Pathway. Oncogene 2018, 37, 1175–1191. [Google Scholar] [CrossRef] [Green Version]
  200. Yu, X.; Zhang, Q.; Zhang, X.; Han, Q.; Li, H.; Mao, Y.; Wang, X.; Guo, H.; Irwin, D.M.; Niu, G.; et al. Exosomes from Macrophages Exposed to Apoptotic Breast Cancer Cells Promote Breast Cancer Proliferation and Metastasis. J. Cancer 2019, 10, 2892–2906. [Google Scholar] [CrossRef] [Green Version]
  201. Xu, Z.; Chen, Y.; Ma, L.; Chen, Y.; Liu, J.; Guo, Y.; Yu, T.; Zhang, L.; Zhu, L.; Shu, Y. Role of Exosomal Non-Coding RNAs from Tumor Cells and Tumor-Associated Macrophages in the Tumor Microenvironment. Mol. Ther. 2022, 30, 3133–3154. [Google Scholar] [CrossRef] [PubMed]
  202. Camera, G.L.; Gelsomino, L.; Malivindi, R.; Barone, I.; Panza, S.; Rose, D.D.; Giordano, F.; D’Esposito, V.; Formisano, P.; Bonofiglio, D.; et al. Adipocyte-Derived Extracellular Vesicles Promote Breast Cancer Cell Malignancy through HIF-1α Activity. Cancer Lett. 2021, 521, 155–168. [Google Scholar] [CrossRef]
  203. Wang, S.; Su, X.; Xu, M.; Xiao, X.; Li, X.; Li, H.; Keating, A.; Zhao, R.C. Exosomes Secreted by Mesenchymal Stromal/Stem Cell-Derived Adipocytes Promote Breast Cancer Cell Growth via Activation of Hippo Signaling Pathway. Stem Cell Res. Ther. 2019, 10, 117. [Google Scholar] [CrossRef] [Green Version]
  204. Ramos-Andrade, I.; Moraes, J.; Brandão-Costa, R.M.; da Silva, S.V.; de Souza, A.; da Silva, C.; Renovato-Martins, M.; Barja-Fidalgo, C. Obese Adipose Tissue Extracellular Vesicles Raise Breast Cancer Cell Malignancy. Endocr.-Relat. Cancer 2020, 27, 571–582. [Google Scholar] [CrossRef] [PubMed]
  205. Lin, R.; Wang, S.; Zhao, R.C. Exosomes from Human Adipose-Derived Mesenchymal Stem Cells Promote Migration through Wnt Signaling Pathway in a Breast Cancer Cell Model. Mol. Cell Biochem. 2013, 383, 13–20. [Google Scholar] [CrossRef] [PubMed]
  206. Khanh, V.C.; Fukushige, M.; Moriguchi, K.; Yamashita, T.; Osaka, M.; Hiramatsu, Y.; Ohneda, O. Type 2 Diabetes Mellitus Induced Paracrine Effects on Breast Cancer Metastasis Through Extracellular Vesicles Derived from Human Mesenchymal Stem Cells. Stem Cells Dev. 2020, 29, 1382–1394. [Google Scholar] [CrossRef]
  207. Wu, S.; Wang, Y.; Yuan, Z.; Wang, S.; Du, H.; Liu, X.; Wang, Q.; Zhu, X. Human Adipose-Derived Mesenchymal Stem Cells Promote Breast Cancer MCF7 Cell Epithelial-Mesenchymal Transition by Cross Interacting with the TGF-β/Smad and PI3K/AKT Signaling Pathways. Mol. Med. Rep. 2019, 19, 177–186. [Google Scholar] [CrossRef] [Green Version]
  208. Moraes, J.A.; Encarnação, C.; Franco, V.A.; Botelho, L.G.X.; Rodrigues, G.P.; Ramos-Andrade, I.; Barja-Fidalgo, C.; Renovato-Martins, M. Adipose Tissue-Derived Extracellular Vesicles and the Tumor Microenvironment: Revisiting the Hallmarks of Cancer. Cancers 2021, 13, 3328. [Google Scholar] [CrossRef]
  209. Mantovani, A.; Garlanda, C.; Allavena, P. Molecular Pathways and Targets in Cancer-Related Inflammation. Ann. Med. 2010, 42, 161–170. [Google Scholar] [CrossRef] [PubMed]
  210. Othman, N.; Jamal, R.; Abu, N. Cancer-Derived Exosomes as Effectors of Key Inflammation-Related Players. Front. Immunol. 2019, 10, 2103. [Google Scholar] [CrossRef] [Green Version]
  211. Chow, A.; Zhou, W.; Liu, L.; Fong, M.Y.; Champer, J.; Haute, D.V.; Chin, A.R.; Ren, X.; Gugiu, B.G.; Meng, Z.; et al. Macrophage Immunomodulation by Breast Cancer-Derived Exosomes Requires Toll-like Receptor 2-Mediated Activation of NF-ΚB. Sci. Rep. 2014, 4, 5750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Hall, J.M.; Zuppan, P.J.; Anderson, L.A.; Huey, B.; Carter, C.; King, M.C. Oncogenes and Human Breast Cancer. Am. J. Hum. Genet. 1989, 44, 577–584. [Google Scholar]
  213. Kalimutho, M.; Nones, K.; Srihari, S.; Duijf, P.H.G.; Waddell, N.; Khanna, K.K. Patterns of Genomic Instability in Breast Cancer. Trends Pharmacol. Sci. 2019, 40, 198–211. [Google Scholar] [CrossRef]
  214. Liu, B.; Wen, X.; Cheng, Y. Survival or Death: Disequilibrating the Oncogenic and Tumor Suppressive Autophagy in Cancer. Cell Death Dis. 2013, 4, e892. [Google Scholar] [CrossRef] [Green Version]
  215. Sanchez-Vega, F.; Mina, M.; Armenia, J.; Chatila, W.K.; Luna, A.; La, K.C.; Dimitriadoy, S.; Liu, D.L.; Kantheti, H.S.; Saghafinia, S.; et al. Oncogenic Signaling Pathways in The Cancer Genome Atlas. Cell 2018, 173, 321–337.e10. [Google Scholar] [CrossRef] [Green Version]
  216. Walerych, D.; Napoli, M.; Collavin, L.; Sal, G.D. The Rebel Angel: Mutant P53 as the Driving Oncogene in Breast Cancer. Carcinogenesis 2012, 33, 2007–2017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Rodríguez, M.; Silva, J.; Herrera, A.; Herrera, M.; Peña, C.; Martín, P.; Gil-Calderón, B.; Larriba, M.J.; Coronado, M.J.; Soldevilla, B.; et al. Exosomes Enriched in Stemness/Metastatic-Related MRNAS Promote Oncogenic Potential in Breast Cancer. Oncotarget 2015, 6, 40575–40587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Xiong, X.; Feng, Y.; Li, L.; Yao, J.; Zhou, M.; Zhao, P.; Huang, F.; Zeng, L.; Yuan, L. Long Non-Coding RNA SNHG1 Promotes Breast Cancer Progression by Regulation of LMO4. Oncol. Rep. 2020, 43, 1503–1515. [Google Scholar] [CrossRef] [Green Version]
  219. Dai, G.; Yang, Y.; Liu, S.; Liu, H. Hypoxic Breast Cancer Cell-Derived Exosomal SNHG1 Promotes Breast Cancer Growth and Angiogenesis via Regulating MiR-216b-5p/JAK2 Axis. Cancer Manag. Res. 2022, 14, 123–133. [Google Scholar] [CrossRef]
  220. Ma, S.; McGuire, M.H.; Mangala, L.S.; Lee, S.; Stur, E.; Hu, W.; Bayraktar, E.; Villar-Prados, A.; Ivan, C.; Wu, S.Y.; et al. Gain-of-Function P53 Protein Transferred via Small Extracellular Vesicles Promotes Conversion of Fibroblasts to a Cancer-Associated Phenotype. Cell Rep. 2021, 34, 108726. [Google Scholar] [CrossRef] [PubMed]
  221. Kilinc, S.; Paisner, R.; Camarda, R.; Gupta, S.; Momcilovic, O.; Kohnz, R.A.; Avsaroglu, B.; L’Etoile, N.D.; Perera, R.M.; Nomura, D.K.; et al. Oncogene-regulated release of extracellular vesicles. Dev Cell. 2021, 56, 1989–2006.e6. [Google Scholar] [CrossRef]
  222. Lehuédé, C.; Dupuy, F.; Rabinovitch, R.; Jones, R.G.; Siegel, P.M. Metabolic Plasticity as a Determinant of Tumor Growth and Metastasis. Cancer Res. 2016, 76, 5201–5208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Ferreira, L.M.R. Cancer Metabolism: The Warburg Effect Today. Exp. Mol. Pathol. 2010, 89, 372–380. [Google Scholar] [CrossRef] [PubMed]
  224. Liberti, M.V.; Locasale, J.W. The Warburg Effect: How Does It Benefit Cancer Cells? Trends Biochem. Sci. 2016, 41, 211–218. [Google Scholar] [CrossRef] [Green Version]
  225. Sciacovelli, M.; Frezza, C. Metabolic Reprogramming and Epithelial-to-mesenchymal Transition in Cancer. FEBS J. 2017, 284, 3132–3144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Lunetti, P.; Giacomo, M.D.; Vergara, D.; Domenico, S.D.; Maffia, M.; Zara, V.; Capobianco, L.; Ferramosca, A. Metabolic Reprogramming in Breast Cancer Results in Distinct Mitochondrial Bioenergetics between Luminal and Basal Subtypes. FEBS J. 2019, 286, 688–709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Vlashi, E.; Lagadec, C.; Vergnes, L.; Reue, K.; Frohnen, P.; Chan, M.; Alhiyari, Y.; Dratver, M.B.; Pajonk, F. Metabolic Differences in Breast Cancer Stem Cells and Differentiated Progeny. Breast Cancer Res. Treat. 2014, 146, 525–534. [Google Scholar] [CrossRef] [Green Version]
  228. Kim, M.-J.; Kim, D.-H.; Jung, W.-H.; Koo, J.-S. Expression of Metabolism-Related Proteins in Triple-Negative Breast Cancer. Int. J. Clin. Exp. Pathol. 2013, 7, 301–312. [Google Scholar]
  229. Evans, K.W.; Yuca, E.; Scott, S.S.; Zhao, M.; Arango, N.P.; Pico, C.X.C.; Saridogan, T.; Shariati, M.; Class, C.A.; Bristow, C.A.; et al. Oxidative Phosphorylation Is a Metabolic Vulnerability in Chemotherapy-Resistant Triple Negative Breast Cancer. Cancer Res. 2021, 81, 5572–5581. [Google Scholar] [CrossRef]
  230. Hirschhaeuser, F.; Sattler, U.G.A.; Mueller-Klieser, W. Lactate: A Metabolic Key Player in Cancer. Cancer Res. 2011, 71, 6921–6925. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Dias, A.S.; Almeida, C.R.; Helguero, L.A.; Duarte, I.F. Metabolic Crosstalk in the Breast Cancer Microenvironment. Eur. J. Cancer 2019, 121, 154–171. [Google Scholar] [CrossRef]
  232. Martinez-Outschoorn, U.E.; Pavlides, S.; Howell, A.; Pestell, R.G.; Tanowitz, H.B.; Sotgia, F.; Lisanti, M.P. Stromal–Epithelial Metabolic Coupling in Cancer: Integrating Autophagy and Metabolism in the Tumor Microenvironment. Int. J. Biochem. Cell Biol. 2011, 43, 1045–1051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Bonuccelli, G.; Whitaker-Menezes, D.; Castello-Cros, R.; Pavlides, S.; Pestell, R.G.; Fatatis, A.; Witkiewicz, A.K.; Heiden, M.G.V.; Migneco, G.; Chiavarina, B.; et al. The Reverse Warburg Effect: Glycolysis Inhibitors Prevent the Tumor Promoting Effects of Caveolin-1 Deficient Cancer Associated Fibroblasts. Cell Cycle 2010, 9, 1960–1971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Zaoui, M.; Morel, M.; Ferrand, N.; Fellahi, S.; Bastard, J.-P.; Lamazière, A.; Larsen, A.K.; Béréziat, V.; Atlan, M.; Sabbah, M. Breast-Associated Adipocytes Secretome Induce Fatty Acid Uptake and Invasiveness in Breast Cancer Cells via CD36 Independently of Body Mass Index, Menopausal Status and Mammary Density. Cancers 2019, 11, 2012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Carracedo, A.; Cantley, L.C.; Pandolfi, P.P. Cancer Metabolism: Fatty Acid Oxidation in the Limelight. Nat. Rev. Cancer 2013, 13, 227–232. [Google Scholar] [CrossRef]
  236. Hensley, C.T.; Wasti, A.T.; DeBerardinis, R.J. Glutamine and Cancer: Cell Biology, Physiology, and Clinical Opportunities. J. Clin. Investig. 2013, 123, 3678–3684. [Google Scholar] [CrossRef] [Green Version]
  237. Risha, Y.; Minic, Z.; Ghobadloo, S.M.; Berezovski, M.V. The Proteomic Analysis of Breast Cell Line Exosomes Reveals Disease Patterns and Potential Biomarkers. Sci. Rep. 2020, 10, 13572. [Google Scholar] [CrossRef]
  238. Joudaki, N.; Rashno, M.; Asadirad, A.; Khodadadi, A. Role of Breast Cancer-Derived Exosomes in Metabolism of Immune Cells through PD1-GLUT1-HK2 Metabolic Axis. Tissue Cell 2021, 71, 101576. [Google Scholar] [CrossRef]
  239. Fong, M.Y.; Zhou, W.; Liu, L.; Alontaga, A.Y.; Chandra, M.; Ashby, J.; Chow, A.; O’Connor, S.T.F.; Li, S.; Chin, A.R.; et al. Breast-Cancer-Secreted MiR-122 Reprograms Glucose Metabolism in Premetastatic Niche to Promote Metastasis. Nat. Cell Biol. 2015, 17, 183–194. [Google Scholar] [CrossRef] [Green Version]
  240. Kang, S.Y.; Lee, E.J.; Byun, J.W.; Han, D.; Choi, Y.; Hwang, D.W.; Lee, D.S. Extracellular Vesicles Induce an Aggressive Phenotype in Luminal Breast Cancer Cells Via PKM2 Phosphorylation. Front. Oncol. 2021, 11, 785450. [Google Scholar] [CrossRef] [PubMed]
  241. Yan, W.; Wu, X.; Zhou, W.; Fong, M.Y.; Cao, M.; Liu, J.; Liu, X.; Chen, C.-H.; Fadare, O.; Pizzo, D.P.; et al. Cancer-Cell-Secreted Exosomal MiR-105 Promotes Tumour Growth through the MYC-Dependent Metabolic Reprogramming of Stromal Cells. Nat. Cell Biol. 2018, 20, 597–609. [Google Scholar] [CrossRef] [Green Version]
  242. Fong, M.Y.; Yan, W.; Ghassemian, M.; Wu, X.; Zhou, X.; Cao, M.; Jiang, L.; Wang, J.; Liu, X.; Zhang, J.; et al. Cancer-secreted MiRNAs Regulate Amino-acid-induced MTORC1 Signaling and Fibroblast Protein Synthesis. EMBO Rep. 2021, 22, e51239. [Google Scholar] [CrossRef] [PubMed]
  243. Wu, Q.; Sun, S.; Li, Z.; Yang, Q.; Li, B.; Zhu, S.; Wang, L.; Wu, J.; Yuan, J.; Yang, C.; et al. Tumour-Originated Exosomal MiR-155 Triggers Cancer-Associated Cachexia to Promote Tumour Progression. Mol. Cancer 2018, 17, 155. [Google Scholar] [CrossRef] [Green Version]
  244. Balaban, S.; Shearer, R.F.; Lee, L.S.; van Geldermalsen, M.; Schreuder, M.; Shtein, H.C.; Cairns, R.; Thomas, K.C.; Fazakerley, D.J.; Grewal, T.; et al. Adipocyte Lipolysis Links Obesity to Breast Cancer Growth: Adipocyte-Derived Fatty Acids Drive Breast Cancer Cell Proliferation and Migration. Cancer Metab. 2017, 5, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Whiteside, T.L. Tumor-Derived Exosomes and Their Role in Cancer Progression. Adv. Clin. Chem. 2016, 74, 103–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Mineo, M.; Garfield, S.H.; Taverna, S.; Flugy, A.; Leo, G.D.; Alessandro, R.; Kohn, E.C. Exosomes Released by K562 Chronic Myeloid Leukemia Cells Promote Angiogenesis in a Src-Dependent Fashion. Angiogenesis 2012, 15, 33–45. [Google Scholar] [CrossRef] [Green Version]
  247. Madu, C.O.; Wang, S.; Madu, C.O.; Lu, Y. Angiogenesis in Breast Cancer Progression, Diagnosis, and Treatment. J. Cancer 2020, 11, 4474–4494. [Google Scholar] [CrossRef]
  248. Akil, A.; Gutiérrez-García, A.K.; Guenter, R.; Rose, J.B.; Beck, A.W.; Chen, H.; Ren, B. Notch Signaling in Vascular Endothelial Cells, Angiogenesis, and Tumor Progression: An Update and Prospective. Front. Cell Dev. Biol. 2021, 9, 642352. [Google Scholar] [CrossRef]
  249. Fagiani, E.; Christofori, G. Angiopoietins in Angiogenesis. Cancer Lett. 2013, 328, 18–26. [Google Scholar] [CrossRef]
  250. Claesson-Welsh, L.; Welsh, M. VEGFA and Tumour Angiogenesis. J. Intern. Med. 2013, 273, 114–127. [Google Scholar] [CrossRef] [Green Version]
  251. King, H.W.; Michael, M.Z.; Gleadle, J.M. Hypoxic Enhancement of Exosome Release by Breast Cancer Cells. BMC Cancer 2012, 12, 421. [Google Scholar] [CrossRef] [Green Version]
  252. Jung, K.O.; Youn, H.; Lee, C.-H.; Kang, K.W.; Chung, J.-K. Visualization of Exosome-Mediated MiR-210 Transfer from Hypoxic Tumor Cells. Oncotarget 2016, 8, 9899–9910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Kosaka, N.; Iguchi, H.; Hagiwara, K.; Yoshioka, Y.; Takeshita, F.; Ochiya, T. Neutral Sphingomyelinase 2 (NSMase2)-Dependent Exosomal Transfer of Angiogenic MicroRNAs Regulate Cancer Cell Metastasis. J. Biol. Chem. 2013, 288, 10849–10859. [Google Scholar] [CrossRef] [Green Version]
  254. Matesanz, N.; Park, G.; McAllister, H.; Leahey, W.; Devine, A.; McVeigh, G.E.; Gardiner, T.A.; McDonald, D.M. Docosahexaenoic Acid Improves the Nitroso-Redox Balance and Reduces VEGF-Mediated Angiogenic Signaling in Microvascular Endothelial Cells. Investig. Ophthalmol. Vis. Sci. 2010, 51, 6815–6825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Ghaffari-Makhmalbaf, P.; Sayyad, M.; Pakravan, K.; Razmara, E.; Bitaraf, A.; Bakhshinejad, B.; Goudarzi, P.; Yousefi, H.; Pournaghshband, M.; Nemati, F.; et al. Docosahexaenoic Acid Reverses the Promoting Effects of Breast Tumor Cell-Derived Exosomes on Endothelial Cell Migration and Angiogenesis. Life Sci. 2021, 264, 118719. [Google Scholar] [CrossRef]
  256. Shi, P.; Liu, Y.; Yang, H.; Hu, B. Breast Cancer Derived Exosomes Promoted Angiogenesis of Endothelial Cells in Microenvironment via CircHIPK3/MiR-124-3p/MTDH Axis. Cell Signal. 2022, 95, 110338. [Google Scholar] [CrossRef] [PubMed]
  257. Feng, Y.; Wang, L.; Wang, T.; Li, Y.; Xun, Q.; Zhang, R.; Liu, L.; Li, L.; Wang, W.; Tian, Y.; et al. Tumor Cell-Secreted Exosomal MiR-22-3p Inhibits Transgelin and Induces Vascular Abnormalization to Promote Tumor Budding. Mol. Ther. 2021, 29, 2151–2166. [Google Scholar] [CrossRef]
  258. Pan, S.; Zhao, X.; Shao, C.; Fu, B.; Huang, Y.; Zhang, N.; Dou, X.; Zhang, Z.; Qiu, Y.; Wang, R.; et al. STIM1 Promotes Angiogenesis by Reducing Exosomal MiR-145 in Breast Cancer MDA-MB-231 Cells. Cell Death Dis. 2021, 12, 38. [Google Scholar] [CrossRef]
  259. Chaudhary, P.; Gibbs, L.D.; Maji, S.; Lewis, C.M.; Suzuki, S.; Vishwanatha, J.K. Serum Exosomal-Annexin A2 Is Associated with African-American Triple-Negative Breast Cancer and Promotes Angiogenesis. Breast Cancer Res. 2020, 22, 11. [Google Scholar] [CrossRef]
  260. Han, B.; Zhang, H.; Tian, R.; Liu, H.; Wang, Z.; Wang, Z.; Tian, J.; Cui, Y.; Ren, S.; Zuo, X.; et al. Exosomal EPHA2 Derived from Highly Metastatic Breast Cancer Cells Promotes Angiogenesis by Activating the AMPK Signaling Pathway through Ephrin A1-EPHA2 Forward Signaling. Theranostics 2022, 12, 4127–4146. [Google Scholar] [CrossRef]
  261. Thompson, C.A.; Purushothaman, A.; Ramani, V.C.; Vlodavsky, I.; Sanderson, R.D. Heparanase Regulates Secretion, Composition, and Function of Tumor Cell-Derived Exosomes. J. Biol. Chem. 2013, 288, 10093–10099. [Google Scholar] [CrossRef] [Green Version]
  262. Lee, J.-K.; Park, S.-R.; Jung, B.-K.; Jeon, Y.-K.; Lee, Y.-S.; Kim, M.-K.; Kim, Y.-G.; Jang, J.-Y.; Kim, C.-W. Exosomes Derived from Mesenchymal Stem Cells Suppress Angiogenesis by Down-Regulating VEGF Expression in Breast Cancer Cells. PLoS ONE 2013, 8, e84256. [Google Scholar] [CrossRef] [Green Version]
  263. Pakravan, K.; Babashah, S.; Sadeghizadeh, M.; Mowla, S.J.; Mossahebi-Mohammadi, M.; Ataei, F.; Dana, N.; Javan, M. MicroRNA-100 Shuttled by Mesenchymal Stem Cell-Derived Exosomes Suppresses in Vitro Angiogenesis through Modulating the MTOR/HIF-1α/VEGF Signaling Axis in Breast Cancer Cells. Cell Oncol. 2017, 40, 457–470. [Google Scholar] [CrossRef] [PubMed]
  264. Schreiber, R.D.; Old, L.J.; Smyth, M.J. Cancer Immunoediting: Integrating Immunity’s Roles in Cancer Suppression and Promotion. Science 2011, 331, 1565–1570. [Google Scholar] [CrossRef] [Green Version]
  265. Alcazar, C.R.G.D.; Alečković, M.; Polyak, K. Immune Escape during Breast Tumor Progression. Cancer Immunol. Res. 2020, 8, 422–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Spranger, S.; Sivan, A.; Corrales, L.; Gajewski, T.F. Tumor and Host Factors Controlling Antitumor Immunity and Efficacy of Cancer Immunotherapy. Adv. Immunol. 2016, 130, 75–93. [Google Scholar] [CrossRef] [Green Version]
  267. Angelova, M.; Mlecnik, B.; Vasaturo, A.; Bindea, G.; Fredriksen, T.; Lafontaine, L.; Buttard, B.; Morgand, E.; Bruni, D.; Jouret-Mourin, A.; et al. Evolution of Metastases in Space and Time under Immune Selection. Cell 2018, 175, 751–765.e16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Steven, A.; Seliger, B. The Role of Immune Escape and Immune Cell Infiltration in Breast Cancer. Breast Care 2018, 13, 16–21. [Google Scholar] [CrossRef] [PubMed]
  269. Gatti-Mays, M.E.; Balko, J.M.; Gameiro, S.R.; Bear, H.D.; Prabhakaran, S.; Fukui, J.; Disis, M.L.; Nanda, R.; Gulley, J.L.; Kalinsky, K.; et al. If We Build It They Will Come: Targeting the Immune Response to Breast Cancer. NPJ Breast Cancer 2019, 5, 37. [Google Scholar] [CrossRef] [Green Version]
  270. Record, M.; Subra, C.; Silvente-Poirot, S.; Poirot, M. Exosomes as Intercellular Signalosomes and Pharmacological Effectors. Biochem. Pharmacol. 2011, 81, 1171–1182. [Google Scholar] [CrossRef] [Green Version]
  271. Whiteside, T.L. Exosomes in Cancer: Another Mechanism of Tumor-Induced Immune Suppression. Adv. Exp. Med. Biol. 2017, 1036, 81–89. [Google Scholar] [CrossRef]
  272. Whiteside, T.L. Exosomes Carrying Immunoinhibitory Proteins and Their Role in Cancer. Clin. Amp. Exp. Immunol. 2017, 189, 259–267. [Google Scholar] [CrossRef] [Green Version]
  273. Jiang, Y.; Li, Y.; Zhu, B. T-Cell Exhaustion in the Tumor Microenvironment. Cell Death Dis. 2015, 6, e1792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. RONG, L.; LI, R.; LI, S.; LUO, R. Immunosuppression of Breast Cancer Cells Mediated by Transforming Growth Factor-β in Exosomes from Cancer Cells. Oncol. Lett. 2016, 11, 500–504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Wieckowski, E.U.; Visus, C.; Szajnik, M.; Szczepanski, M.J.; Storkus, W.J.; Whiteside, T.L. Tumor-Derived Microvesicles Promote Regulatory T Cell Expansion and Induce Apoptosis in Tumor-Reactive Activated CD8+ T Lymphocytes. J. Immunol. 2009, 183, 3720–3730. [Google Scholar] [CrossRef] [Green Version]
  276. Ning, Y.; Shen, K.; Wu, Q.; Sun, X.; Bai, Y.; Xie, Y.; Pan, J.; Qi, C. Tumor Exosomes Block Dendritic Cells Maturation to Decrease the T Cell Immune Response. Immunol. Lett. 2018, 199, 36–43. [Google Scholar] [CrossRef]
  277. Ni, C.; Fang, Q.-Q.; Chen, W.-Z.; Jiang, J.-X.; Jiang, Z.; Ye, J.; Zhang, T.; Yang, L.; Meng, F.-B.; Xia, W.-J.; et al. Breast Cancer-Derived Exosomes Transmit LncRNA SNHG16 to Induce CD73+γδ1 Treg Cells. Signal. Transduct. Target. Ther. 2020, 5, 41. [Google Scholar] [CrossRef]
  278. Piao, Y.J.; Kim, H.S.; Hwang, E.H.; Woo, J.; Zhang, M.; Moon, W.K. Breast Cancer Cell-Derived Exosomes and Macrophage Polarization Are Associated with Lymph Node Metastasis. Oncotarget 2017, 9, 7398–7410. [Google Scholar] [CrossRef] [Green Version]
  279. Weng, Y.-S.; Tseng, H.-Y.; Chen, Y.-A.; Shen, P.-C.; Haq, A.T.A.; Chen, L.-M.; Tung, Y.-C.; Hsu, H.-L. MCT-1/MiR-34a/IL-6/IL-6R Signaling Axis Promotes EMT Progression, Cancer Stemness and M2 Macrophage Polarization in Triple-Negative Breast Cancer. Mol. Cancer 2019, 18, 42. [Google Scholar] [CrossRef] [Green Version]
  280. Xing, F.; Liu, Y.; Wu, S.-Y.; Wu, K.; Sharma, S.; Mo, Y.-Y.; Feng, J.; Sanders, S.; Jin, G.; Singh, R.; et al. Correction: Loss of XIST in Breast Cancer Activates MSN-c-Met and Reprograms Microglia via Exosomal MiRNA to Promote Brain Metastasis. Cancer Res. 2021, 81, 5582. [Google Scholar] [CrossRef] [PubMed]
  281. Mohapatra, S.; Pioppini, C.; Ozpolat, B.; Calin, G.A. Non-Coding RNAs Regulation of Macrophage Polarization in Cancer. Mol. Cancer 2021, 20, 24. [Google Scholar] [CrossRef]
  282. Dou, D.; Ren, X.; Han, M.; Xu, X.; Ge, X.; Gu, Y.; Wang, X. Cancer-Associated Fibroblasts-Derived Exosomes Suppress Immune Cell Function in Breast Cancer via the MiR-92/PD-L1 Pathway. Front. Immunol. 2020, 11, 2026. [Google Scholar] [CrossRef] [PubMed]
  283. Qi, M.; Xia, Y.; Wu, Y.; Zhang, Z.; Wang, X.; Lu, L.; Dai, C.; Song, Y.; Xu, K.; Ji, W.; et al. Lin28B-High Breast Cancer Cells Promote Immune Suppression in the Lung Pre-Metastatic Niche via Exosomes and Support Cancer Progression. Nat. Commun. 2022, 13, 897. [Google Scholar] [CrossRef] [PubMed]
  284. Liu, C.; Yu, S.; Zinn, K.; Wang, J.; Zhang, L.; Jia, Y.; Kappes, J.C.; Barnes, S.; Kimberly, R.P.; Grizzle, W.E.; et al. Murine Mammary Carcinoma Exosomes Promote Tumor Growth by Suppression of NK Cell Function. J. Immunol. 2006, 176, 1375–1385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Zhuyan, J.; Chen, M.; Zhu, T.; Bao, X.; Zhen, T.; Xing, K.; Wang, Q.; Zhu, S. Critical Steps to Tumor Metastasis: Alterations of Tumor Microenvironment and Extracellular Matrix in the Formation of Pre-Metastatic and Metastatic Niche. Cell Biosci. 2020, 10, 89. [Google Scholar] [CrossRef]
  286. Mani, S.A.; Guo, W.; Liao, M.-J.; Eaton, E.N.; Ayyanan, A.; Zhou, A.Y.; Brooks, M.; Reinhard, F.; Zhang, C.C.; Shipitsin, M.; et al. The Epithelial-Mesenchymal Transition Generates Cells with Properties of Stem Cells. Cell 2008, 133, 704–715. [Google Scholar] [CrossRef] [Green Version]
  287. Yu, M.; Bardia, A.; Wittner, B.S.; Stott, S.L.; Smas, M.E.; Ting, D.T.; Isakoff, S.J.; Ciciliano, J.C.; Wells, M.N.; Shah, A.M.; et al. Circulating Breast Tumor Cells Exhibit Dynamic Changes in Epithelial and Mesenchymal Composition. Science 2013, 339, 580–584. [Google Scholar] [CrossRef] [Green Version]
  288. Vella, L. The Emerging Role of Exosomes in Epithelial–Mesenchymal-Transition in Cancer. Front. Oncol. 2014, 4, 361. [Google Scholar] [CrossRef] [Green Version]
  289. Kletukhina, S.; Neustroeva, O.; James, V.; Rizvanov, A.; Gomzikova, M. Role of Mesenchymal Stem Cell-Derived Extracellular Vesicles in Epithelial–Mesenchymal Transition. Int. J. Mol. Sci. 2019, 20, 4813. [Google Scholar] [CrossRef] [Green Version]
  290. Brena, D.; Huang, M.-B.; Bond, V. Extracellular Vesicle-Mediated Transport: Reprogramming a Tumor Microenvironment Conducive with Breast Cancer Progression and Metastasis. Transl. Oncol. 2022, 15, 101286. [Google Scholar] [CrossRef]
  291. Jiang, J.; Li, J.; Zhou, X.; Zhao, X.; Huang, B.; Qin, Y. Exosomes Regulate the Epithelial–Mesenchymal Transition in Cancer. Front. Oncol. 2022, 12, 864980. [Google Scholar] [CrossRef]
  292. Gwak, J.M.; Kim, H.J.; Kim, E.J.; Chung, Y.R.; Yun, S.; Seo, A.N.; Lee, H.J.; Park, S.Y. MicroRNA-9 Is Associated with Epithelial-Mesenchymal Transition, Breast Cancer Stem Cell Phenotype, and Tumor Progression in Breast Cancer. Breast Cancer Res. Treat. 2014, 147, 39–49. [Google Scholar] [CrossRef] [PubMed]
  293. Drasin, D.J.; Guarnieri, A.L.; Neelakantan, D.; Kim, J.; Cabrera, J.H.; Wang, C.-A.; Zaberezhnyy, V.; Gasparini, P.; Cascione, L.; Huebner, K.; et al. TWIST1-Induced MiR-424 Reversibly Drives Mesenchymal Programming While Inhibiting Tumor Initiation. Cancer Res. 2015, 75, 1908–1921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Green, T.M.; Alpaugh, M.L.; Barsky, S.H.; Rappa, G.; Lorico, A. Breast Cancer-Derived Extracellular Vesicles: Characterization and Contribution to the Metastatic Phenotype. Biomed. Res. Int. 2015, 2015, 634865. [Google Scholar] [CrossRef] [Green Version]
  295. Zhao, Y.; Zheng, X.; Zheng, Y.; Chen, Y.; Fei, W.; Wang, F.; Zheng, C. Extracellular Matrix: Emerging Roles and Potential Therapeutic Targets for Breast Cancer. Front. Oncol. 2021, 11, 650453. [Google Scholar] [CrossRef]
  296. Das, A.; Mohan, V.; Krishnaswamy, V.R.; Solomonov, I.; Sagi, I. Exosomes as a Storehouse of Tissue Remodeling Proteases and Mediators of Cancer Progression. Cancer Metastasis Rev. 2019, 38, 455–468. [Google Scholar] [CrossRef] [PubMed]
  297. Winkler, J.; Abisoye-Ogunniyan, A.; Metcalf, K.J.; Werb, Z. Concepts of Extracellular Matrix Remodelling in Tumour Progression and Metastasis. Nat. Commun. 2020, 11, 5120. [Google Scholar] [CrossRef]
  298. Han, W.; Chen, S.; Yuan, W.; Fan, Q.; Tian, J.; Wang, X.; Chen, L.; Zhang, X.; Wei, W.; Liu, R.; et al. Oriented Collagen Fibers Direct Tumor Cell Intravasation. Proc. Natl. Acad. Sci. USA 2016, 113, 11208–11213. [Google Scholar] [CrossRef] [Green Version]
  299. Patwardhan, S.; Mahadik, P.; Shetty, O.; Sen, S. ECM Stiffness-Tuned Exosomes Drive Breast Cancer Motility through Thrombospondin-1. Biomaterials 2021, 279, 121185. [Google Scholar] [CrossRef]
  300. Wang, J.; Zhang, Q.; Wang, D.; Yang, S.; Zhou, S.; Xu, H.; Zhang, H.; Zhong, S.; Feng, J. Microenvironment-induced TIMP2 Loss by Cancer-secreted Exosomal MiR-4443 Promotes Liver Metastasis of Breast Cancer. J. Cell Physiol. 2020, 235, 5722–5735. [Google Scholar] [CrossRef]
  301. Gupta, G.P.; Nguyen, D.X.; Chiang, A.C.; Bos, P.D.; Kim, J.Y.; Nadal, C.; Gomis, R.R.; Manova-Todorova, K.; Massagué, J. Mediators of Vascular Remodelling Co-Opted for Sequential Steps in Lung Metastasis. Nature 2007, 446, 765–770. [Google Scholar] [CrossRef]
  302. Ghoroghi, S.; Mary, B.; Larnicol, A.; Asokan, N.; Klein, A.; Osmani, N.; Busnelli, I.; Delalande, F.; Paul, N.; Halary, S.; et al. Ral GTPases Promote Breast Cancer Metastasis by Controlling Biogenesis and Organ Targeting of Exosomes. Elife 2021, 10, e61539. [Google Scholar] [CrossRef] [PubMed]
  303. Oskarsson, T.; Batlle, E.; Massagué, J. Metastatic Stem Cells: Sources, Niches, and Vital Pathways. Cell Stem Cell 2014, 14, 306–321. [Google Scholar] [CrossRef] [Green Version]
  304. Terceiro, L.E.L.; Edechi, C.A.; Ikeogu, N.M.; Nickel, B.E.; Hombach-Klonisch, S.; Sharif, T.; Leygue, E.; Myal, Y. The Breast Tumor Microenvironment: A Key Player in Metastatic Spread. Cancers 2021, 13, 4798. [Google Scholar] [CrossRef] [PubMed]
  305. Wang, X.; Sun, C.; Huang, X.; Li, J.; Fu, Z.; Li, W.; Yin, Y. The Advancing Roles of Exosomes in Breast Cancer. Front. Cell Dev. Biol. 2021, 9, 731062. [Google Scholar] [CrossRef] [PubMed]
  306. Luga, V.; Zhang, L.; Viloria-Petit, A.M.; Ogunjimi, A.A.; Inanlou, M.R.; Chiu, E.; Buchanan, M.; Hosein, A.N.; Basik, M.; Wrana, J.L. Exosomes Mediate Stromal Mobilization of Autocrine Wnt-PCP Signaling in Breast Cancer Cell Migration. Cell 2012, 151, 1542–1556. [Google Scholar] [CrossRef] [Green Version]
  307. Shimoda, M.; Principe, S.; Jackson, H.W.; Luga, V.; Fang, H.; Molyneux, S.D.; Shao, Y.W.; Aiken, A.; Waterhouse, P.D.; Karamboulas, C.; et al. Loss of the Timp Gene Family Is Sufficient for the Acquisition of the CAF-like Cell State. Nat. Cell Biol. 2014, 16, 889–901. [Google Scholar] [CrossRef]
  308. Liu, Y.; Xiang, X.; Zhuang, X.; Zhang, S.; Liu, C.; Cheng, Z.; Michalek, S.; Grizzle, W.; Zhang, H.-G. Contribution of MyD88 to the Tumor Exosome-Mediated Induction of Myeloid Derived Suppressor Cells. Am. J. Pathol. 2010, 176, 2490–2499. [Google Scholar] [CrossRef]
  309. Chang, J.; Chaudhuri, O. Beyond Proteases: Basement Membrane Mechanics and Cancer Invasion. J. Cell Biol. 2019, 218, 2456–2469. [Google Scholar] [CrossRef] [Green Version]
  310. Condeelis, J.; Segall, J.E. Intravital Imaging of Cell Movement in Tumours. Nat. Rev. Cancer 2003, 3, 921–930. [Google Scholar] [CrossRef]
  311. Modica, M.D.; Regondi, V.; Sandri, M.; Iorio, M.V.; Zanetti, A.; Tagliabue, E.; Casalini, P.; Triulzi, T. Breast Cancer-Secreted MiR-939 Downregulates VE-Cadherin and Destroys the Barrier Function of Endothelial Monolayers. Cancer Lett. 2017, 384, 94–100. [Google Scholar] [CrossRef] [Green Version]
  312. Tominaga, N.; Kosaka, N.; Ono, M.; Katsuda, T.; Yoshioka, Y.; Tamura, K.; Lötvall, J.; Nakagama, H.; Ochiya, T. Brain Metastatic Cancer Cells Release MicroRNA-181c-Containing Extracellular Vesicles Capable of Destructing Blood–Brain Barrier. Nat. Commun. 2015, 6, 6716. [Google Scholar] [CrossRef] [Green Version]
  313. Peinado, H.; Zhang, H.; Matei, I.R.; Costa-Silva, B.; Hoshino, A.; Rodrigues, G.; Psaila, B.; Kaplan, R.N.; Bromberg, J.F.; Kang, Y.; et al. Pre-Metastatic Niches: Organ-Specific Homes for Metastases. Nat. Rev. Cancer 2017, 17, 302–317. [Google Scholar] [CrossRef] [PubMed]
  314. Wang, H.-X.; Olivier, G. Tumor-Derived Extracellular Vesicles in Breast Cancer: From Bench to Bedside. Cancer Lett. 2019, 460, 54–64. [Google Scholar] [CrossRef] [PubMed]
  315. Hashimoto, K.; Ochi, H.; Sunamura, S.; Kosaka, N.; Mabuchi, Y.; Fukuda, T.; Yao, K.; Kanda, H.; Ae, K.; Okawa, A.; et al. Cancer-Secreted Hsa-MiR-940 Induces an Osteoblastic Phenotype in the Bone Metastatic Microenvironment via Targeting ARHGAP1 and FAM134A. Proc. Natl. Acad. Sci. USA 2018, 115, 2204–2209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Akhtar, M.; Haider, A.; Rashid, S.; Al-Nabet, A.D.M.H. Paget’s “Seed and Soil” Theory of Cancer Metastasis: An Idea Whose Time Has Come. Adv. Anat. Pathol. 2019, 26, 69–74. [Google Scholar] [CrossRef]
  317. Chin, A.R.; Wang, S.E. Cancer-Derived Extracellular Vesicles: The ‘Soil Conditioner’ in Breast Cancer Metastasis? Cancer Metast. Rev 2016, 35, 669–676. [Google Scholar] [CrossRef]
  318. Liu, Q.; Zhang, H.; Jiang, X.; Qian, C.; Liu, Z.; Luo, D. Factors Involved in Cancer Metastasis: A Better Understanding to “Seed and Soil” Hypothesis. Mol. Cancer 2017, 16, 176. [Google Scholar] [CrossRef] [Green Version]
  319. Cox, T.R.; Rumney, R.M.H.; Schoof, E.M.; Perryman, L.; Høye, A.M.; Agrawal, A.; Bird, D.; Latif, N.A.; Forrest, H.; Evans, H.R.; et al. The Hypoxic Cancer Secretome Induces Pre-Metastatic Bone Lesions through Lysyl Oxidase. Nature 2015, 522, 106–110. [Google Scholar] [CrossRef] [Green Version]
  320. Hoshino, A.; Costa-Silva, B.; Shen, T.-L.; Rodrigues, G.; Hashimoto, A.; Mark, M.T.; Molina, H.; Kohsaka, S.; Giannatale, A.D.; Ceder, S.; et al. Tumour Exosome Integrins Determine Organotropic Metastasis. Nature 2015, 527, 329–335. [Google Scholar] [CrossRef] [Green Version]
  321. Loftus, A.; Cappariello, A.; George, C.; Ucci, A.; Shefferd, K.; Green, A.; Paone, R.; Ponzetti, M.; Monache, S.D.; Muraca, M.; et al. Extracellular Vesicles From Osteotropic Breast Cancer Cells Affect Bone Resident Cells. J. Bone Miner. Res. 2020, 35, 396–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  322. Rodrigues, G.; Hoshino, A.; Kenific, C.M.; Matei, I.R.; Steiner, L.; Freitas, D.; Kim, H.S.; Oxley, P.R.; Scandariato, I.; Casanova-Salas, I.; et al. Tumour Exosomal CEMIP Protein Promotes Cancer Cell Colonization in Brain Metastasis. Nat. Cell Biol. 2019, 21, 1403–1412. [Google Scholar] [CrossRef]
  323. Zhang, L.; Zhang, S.; Yao, J.; Lowery, F.J.; Zhang, Q.; Huang, W.-C.; Li, P.; Li, M.; Wang, X.; Zhang, C.; et al. Microenvironment-Induced PTEN Loss by Exosomal MicroRNA Primes Brain Metastasis Outgrowth. Nature 2015, 527, 100–104. [Google Scholar] [CrossRef] [Green Version]
  324. Yang, X.; Zhang, Y.; Zhang, Y.; Zhang, S.; Qiu, L.; Zhuang, Z.; Wei, M.; Deng, X.; Wang, Z.; Han, J. The Key Role of Exosomes on the Pre-Metastatic Niche Formation in Tumors. Front. Mol. Biosci. 2021, 8, 703640. [Google Scholar] [CrossRef] [PubMed]
  325. Wen, S.W.; Sceneay, J.; Lima, L.G.; Wong, C.S.F.; Becker, M.; Krumeich, S.; Lobb, R.J.; Castillo, V.; Wong, K.N.; Ellis, S.; et al. The Biodistribution and Immune Suppressive Effects of Breast Cancer–Derived Exosomes. Cancer Res. 2016, 76, 6816–6827. [Google Scholar] [CrossRef] [Green Version]
  326. Zhang, H.; Yu, Y.; Zhou, L.; Ma, J.; Tang, K.; Xu, P.; Ji, T.; Liang, X.; Lv, J.; Dong, W.; et al. Circulating Tumor Microparticles Promote Lung Metastasis by Reprogramming Inflammatory and Mechanical Niches via a Macrophage-Dependent Pathway. Cancer Immunol. Res. 2018, 6, 1046–1056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  327. Cheng, X.; Wang, Z. Immune Modulation of Metastatic Niche Formation in the Bone. Front. Immunol. 2021, 12, 765994. [Google Scholar] [CrossRef] [PubMed]
  328. Keklikoglou, I.; Cianciaruso, C.; Güç, E.; Squadrito, M.L.; Spring, L.M.; Tazzyman, S.; Lambein, L.; Poissonnier, A.; Ferraro, G.B.; Baer, C.; et al. Chemotherapy Elicits Pro-Metastatic Extracellular Vesicles in Breast Cancer Models. Nat. Cell Biol. 2019, 21, 190–202. [Google Scholar] [CrossRef] [Green Version]
  329. Qian, B.; Deng, Y.; Im, J.H.; Muschel, R.J.; Zou, Y.; Li, J.; Lang, R.A.; Pollard, J.W. A Distinct Macrophage Population Mediates Metastatic Breast Cancer Cell Extravasation, Establishment and Growth. PLoS ONE 2009, 4, e6562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  330. Ma, R.-Y.; Zhang, H.; Li, X.-F.; Zhang, C.-B.; Selli, C.; Tagliavini, G.; Lam, A.D.; Prost, S.; Sims, A.H.; Hu, H.-Y.; et al. Monocyte-Derived Macrophages Promote Breast Cancer Bone Metastasis Outgrowth. J. Exp. Med. 2020, 217, e20191820. [Google Scholar] [CrossRef]
  331. Qian, B.-Z.; Zhang, H.; Li, J.; He, T.; Yeo, E.-J.; Soong, D.Y.H.; Carragher, N.O.; Munro, A.; Chang, A.; Bresnick, A.R.; et al. FLT1 Signaling in Metastasis-Associated Macrophages Activates an Inflammatory Signature That Promotes Breast Cancer Metastasis. J. Cell Biol. 2015, 210, 2104OIA168. [Google Scholar] [CrossRef] [Green Version]
  332. Banys, M.; Hartkopf, A.D.; Krawczyk, N.; Kaiser, T.; Meier-Stiegen, F.; Fehm, T.; Neubauer, H. Dormancy in Breast Cancer. Breast Cancer Dove Med. Press 2012, 4, 183–191. [Google Scholar] [CrossRef] [Green Version]
  333. Klein, C.A. Framework Models of Tumor Dormancy from Patient-Derived Observations. Curr. Opin. Genet. Dev. 2011, 21, 42–49. [Google Scholar] [CrossRef]
  334. Bushnell, G.G.; Deshmukh, A.P.; den Hollander, P.; Luo, M.; Soundararajan, R.; Jia, D.; Levine, H.; Mani, S.A.; Wicha, M.S. Breast Cancer Dormancy: Need for Clinically Relevant Models to Address Current Gaps in Knowledge. NPJ Breast Cancer 2021, 7, 66. [Google Scholar] [CrossRef] [PubMed]
  335. Angelis, M.L.D.; Francescangeli, F.; Torre, F.L.; Zeuner, A. Stem Cell Plasticity and Dormancy in the Development of Cancer Therapy Resistance. Front. Oncol. 2019, 9, 626. [Google Scholar] [CrossRef] [Green Version]
  336. Angelis, M.L.D.; Francescangeli, F.; Zeuner, A. Breast Cancer Stem Cells as Drivers of Tumor Chemoresistance, Dormancy and Relapse: New Challenges and Therapeutic Opportunities. Cancers 2019, 11, 1569. [Google Scholar] [CrossRef] [Green Version]
  337. Ren, Q.; Khoo, W.H.; Corr, A.P.; Phan, T.G.; Croucher, P.I.; Stewart, S.A. Gene Expression Predicts Dormant Metastatic Breast Cancer Cell Phenotype. Breast Cancer Res. 2022, 24, 10. [Google Scholar] [CrossRef]
  338. Muscarella, A.M.; Aguirre, S.; Hao, X.; Waldvogel, S.M.; Zhang, X.H.-F. Exploiting Bone Niches: Progression of Disseminated Tumor Cells to Metastasis. J. Clin. Investig. 2021, 131, e143764. [Google Scholar] [CrossRef] [PubMed]
  339. Ghajar, C.M. Metastasis Prevention by Targeting the Dormant Niche. Nat. Rev. Cancer 2015, 15, 238–247. [Google Scholar] [CrossRef] [PubMed]
  340. Jahangiri, L.; Ishola, T. Dormancy in Breast Cancer, the Role of Autophagy, LncRNAs, MiRNAs and Exosomes. Int. J. Mol. Sci. 2022, 23, 5271. [Google Scholar] [CrossRef] [PubMed]
  341. Lim, P.K.; Bliss, S.A.; Patel, S.A.; Taborga, M.; Dave, M.A.; Gregory, L.A.; Greco, S.J.; Bryan, M.; Patel, P.S.; Rameshwar, P. Gap Junction–Mediated Import of MicroRNA from Bone Marrow Stromal Cells Can Elicit Cell Cycle Quiescence in Breast Cancer Cells. Cancer Res. 2011, 71, 1550–1560. [Google Scholar] [CrossRef] [Green Version]
  342. Lim, H.-Y.; Wang, W.; Wessells, R.J.; Ocorr, K.; Bodmer, R. Phospholipid Homeostasis Regulates Lipid Metabolism and Cardiac Function through SREBP Signaling in Drosophila. Genes Dev. 2011, 25, 189–200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Shupp, A.B.; Neupane, M.; Agostini, L.C.; Ning, G.; Brody, J.R.; Bussard, K.M. Stromal-Derived Extracellular Vesicles Suppress Proliferation of Bone Metastatic Cancer Cells Mediated By ERK2. Mol. Cancer Res. 2021, 19, 1763–1777. [Google Scholar] [CrossRef] [PubMed]
  344. Ono, M.; Kosaka, N.; Tominaga, N.; Yoshioka, Y.; Takeshita, F.; Takahashi, R.; Yoshida, M.; Tsuda, H.; Tamura, K.; Ochiya, T. Exosomes from Bone Marrow Mesenchymal Stem Cells Contain a MicroRNA That Promotes Dormancy in Metastatic Breast Cancer Cells. Sci. Signal. 2014, 7, ra63. [Google Scholar] [CrossRef]
  345. Vallabhaneni, K.C.; Penfornis, P.; Xing, F.; Hassler, Y.; Adams, K.V.; Mo, Y.-Y.; Watabe, K.; Pochampally, R. Stromal Cell Extracellular Vesicular Cargo Mediated Regulation of Breast Cancer Cell Metastasis via Ubiquitin Conjugating Enzyme E2 N Pathway. Oncotarget 2017, 8, 109861–109876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  346. Ali, N.M.; Yeap, S.K.; Ho, W.Y.; Boo, L.; Ky, H.; Satharasinghe, D.A.; Tan, S.W.; Cheong, S.K.; Huang, H.D.; Lan, K.C.; et al. Adipose MSCs Suppress MCF7 and MDA-MB-231 Breast Cancer Metastasis and EMT Pathways Leading to Dormancy via Exosomal-MiRNAs Following Co-Culture Interaction. Pharmaceuticals 2020, 14, 8. [Google Scholar] [CrossRef]
  347. Walker, N.D.; Elias, M.; Guiro, K.; Bhatia, R.; Greco, S.J.; Bryan, M.; Gergues, M.; Sandiford, O.A.; Ponzio, N.M.; Leibovich, S.J.; et al. Exosomes from Differentially Activated Macrophages Influence Dormancy or Resurgence of Breast Cancer Cells within Bone Marrow Stroma. Cell Death Dis. 2019, 10, 59. [Google Scholar] [CrossRef] [Green Version]
  348. DeSantis, C.E.; Ma, J.; Gaudet, M.M.; Newman, L.A.; Miller, K.D.; Sauer, A.G.; Jemal, A.; Siegel, R.L. Breast Cancer Statistics, 2019. CA Cancer J. Clin. 2019, 69, 438–451. [Google Scholar] [CrossRef] [Green Version]
  349. Nilsson, S.; Mäkelä, S.; Treuter, E.; Tujague, M.; Thomsen, J.; Andersson, G.; Enmark, E.; Pettersson, K.; Warner, M.; Gustafsson, J.-Å. Mechanisms of Estrogen Action. Physiol. Rev. 2001, 81, 1535–1565. [Google Scholar] [CrossRef]
  350. Miller, K.D.; Nogueira, L.; Devasia, T.; Mariotto, A.B.; Yabroff, K.R.; Jemal, A.; Kramer, J.; Siegel, R.L. Cancer Treatment and Survivorship Statistics, 2022. CA Cancer J. Clin. 2022, 72, 409–436. [Google Scholar] [CrossRef] [PubMed]
  351. Osborne, C.K.; Schiff, R. Mechanisms of Endocrine Resistance in Breast Cancer. Annu. Rev. Med. 2011, 62, 233–247. [Google Scholar] [CrossRef] [Green Version]
  352. Clarke, R.; Tyson, J.J.; Dixon, J.M. Endocrine Resistance in Breast Cancer—An Overview and Update. Mol. Cell Endocrinol. 2015, 418, 220–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  353. AlFakeeh, A.; Brezden-Masley, C. Overcoming Endocrine Resistance in Hormone Receptor-Positive Breast Cancer. Curr. Oncol. 2018, 25, S18–S27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Hanker, A.B.; Sudhan, D.R.; Arteaga, C.L. Overcoming Endocrine Resistance in Breast Cancer. Cancer Cell 2020, 37, 496–513. [Google Scholar] [CrossRef] [PubMed]
  355. Camera, G.L.; Gelsomino, L.; Caruso, A.; Panza, S.; Barone, I.; Bonofiglio, D.; Andò, S.; Giordano, C.; Catalano, S. The Emerging Role of Extracellular Vesicles in Endocrine Resistant Breast Cancer. Cancers 2021, 13, 1160. [Google Scholar] [CrossRef]
  356. Semina, S.E.; Scherbakov, A.M.; Vnukova, A.A.; Bagrov, D.V.; Evtushenko, E.G.; Safronova, V.M.; Golovina, D.A.; Lyubchenko, L.N.; Gudkova, M.V.; Krasil’nikov, M.A. Exosome-Mediated Transfer of Cancer Cell Resistance to Antiestrogen Drugs. Mol. J. Synth. Chem. Nat. Prod. Chem. 2018, 23, 829. [Google Scholar] [CrossRef] [Green Version]
  357. Wei, Y.; Lai, X.; Yu, S.; Chen, S.; Ma, Y.; Zhang, Y.; Li, H.; Zhu, X.; Yao, L.; Zhang, J. Exosomal MiR-221/222 Enhances Tamoxifen Resistance in Recipient ER-Positive Breast Cancer Cells. Breast Cancer Res. Treat. 2014, 147, 423–431. [Google Scholar] [CrossRef]
  358. Liu, J.; Zhu, S.; Tang, W.; Huang, Q.; Mei, Y.; Yang, H. Exosomes from Tamoxifen-Resistant Breast Cancer Cells Transmit Drug Resistance Partly by Delivering MiR-9-5p. Cancer Cell Int. 2021, 21, 55. [Google Scholar] [CrossRef]
  359. Hu, K.; Liu, X.; Li, Y.; Li, Q.; Xu, Y.; Zeng, W.; Zhong, G.; Yu, C. Exosomes Mediated Transfer of Circ_UBE2D2 Enhances the Resistance of Breast Cancer to Tamoxifen by Binding to MiR-200a-3p. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2020, 26, e922253-1–e922253-12. [Google Scholar] [CrossRef] [PubMed]
  360. Gao, Y.; Li, X.; Zeng, C.; Liu, C.; Hao, Q.; Li, W.; Zhang, K.; Zhang, W.; Wang, S.; Zhao, H.; et al. CD63+ Cancer-Associated Fibroblasts Confer Tamoxifen Resistance to Breast Cancer Cells through Exosomal MiR-22. Adv. Sci. 2020, 7, 2002518. [Google Scholar] [CrossRef]
  361. Xu, C.-G.; Yang, M.-F.; Ren, Y.-Q.; Wu, C.-H.; Wang, L.-Q. Exosomes Mediated Transfer of LncRNA UCA1 Results in Increased Tamoxifen Resistance in Breast Cancer Cells. Eur. Rev. Med. Pharmacol. 2016, 20, 4362–4368. [Google Scholar]
  362. Sansone, P.; Berishaj, M.; Rajasekhar, V.K.; Ceccarelli, C.; Chang, Q.; Strillacci, A.; Savini, C.; Shapiro, L.; Bowman, R.L.; Mastroleo, C.; et al. Evolution of Cancer Stem-like Cells in Endocrine-Resistant Metastatic Breast Cancers Is Mediated by Stromal Microvesicles. Cancer Res. 2017, 77, 1927–1941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  363. Augimeri, G.; Camera, G.L.; Gelsomino, L.; Giordano, C.; Panza, S.; Sisci, D.; Morelli, C.; Győrffy, B.; Bonofiglio, D.; Andò, S.; et al. Evidence for Enhanced Exosome Production in Aromatase Inhibitor-Resistant Breast Cancer Cells. Int. J. Mol. Sci. 2020, 21, 5841. [Google Scholar] [CrossRef]
  364. Gradishar, W.J.; Anderson, B.O.; Abraham, J.; Aft, R.; Agnese, D.; Allison, K.H.; Blair, S.L.; Burstein, H.J.; Dang, C.; Elias, A.D.; et al. Breast Cancer, Version 3.2020, NCCN Clinical Practice Guidelines in Oncology. J. Natl. Compr. Cancer Netw. 2020, 18, 452–478. [Google Scholar] [CrossRef] [Green Version]
  365. Shedden, K.; Xie, X.T.; Chandaroy, P.; Chang, Y.T.; Rosania, G.R. Expulsion of Small Molecules in Vesicles Shed by Cancer Cells: Association with Gene Expression and Chemosensitivity Profiles. Cancer Res. 2003, 63, 4331–4337. [Google Scholar] [PubMed]
  366. Ifergan, I.; Scheffer, G.L.; Assaraf, Y.G. Novel Extracellular Vesicles Mediate an ABCG2-Dependent Anticancer Drug Sequestration and Resistance. Cancer Res. 2005, 65, 10952–10958. [Google Scholar] [CrossRef] [Green Version]
  367. Lv, M.; Zhu, X.; Chen, W.; Zhong, S.; Hu, Q.; Ma, T.; Zhang, J.; Chen, L.; Tang, J.; Zhao, J. Exosomes Mediate Drug Resistance Transfer in MCF-7 Breast Cancer Cells and a Probable Mechanism Is Delivery of P-Glycoprotein. Tumor Biol. 2014, 35, 10773–10779. [Google Scholar] [CrossRef]
  368. Ma, X.; Cai, Y.; He, D.; Zou, C.; Zhang, P.; Lo, C.Y.; Xu, Z.; Chan, F.L.; Yu, S.; Chen, Y.; et al. Transient Receptor Potential Channel TRPC5 Is Essential for P-Glycoprotein Induction in Drug-Resistant Cancer Cells. Proc. Natl. Acad. Sci. USA 2012, 109, 16282–16287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  369. Dong, Y.; Pan, Q.; Jiang, L.; Chen, Z.; Zhang, F.; Liu, Y.; Xing, H.; Shi, M.; Li, J.; Li, X.; et al. Tumor Endothelial Expression of P-Glycoprotein upon Microvesicular Transfer of TrpC5 Derived from Adriamycin-Resistant Breast Cancer Cells. Biochem. Biophys. Res. Commun. 2014, 446, 85–90. [Google Scholar] [CrossRef] [PubMed]
  370. Wang, T.; Ning, K.; Lu, T.; Sun, X.; Jin, L.; Qi, X.; Jin, J.; Hua, D. Increasing Circulating Exosomes-carrying TRPC5 Predicts Chemoresistance in Metastatic Breast Cancer Patients. Cancer Sci. 2017, 108, 448–454. [Google Scholar] [CrossRef] [Green Version]
  371. Ma, X.; Chen, Z.; Hua, D.; He, D.; Wang, L.; Zhang, P.; Wang, J.; Cai, Y.; Gao, C.; Zhang, X.; et al. Essential Role for TrpC5-Containing Extracellular Vesicles in Breast Cancer with Chemotherapeutic Resistance. Proc. Natl. Acad. Sci. USA 2014, 111, 6389–6394. [Google Scholar] [CrossRef] [Green Version]
  372. Ning, K.; Wang, T.; Sun, X.; Zhang, P.; Chen, Y.; Jin, J.; Hua, D. UCH-L1-containing Exosomes Mediate Chemotherapeutic Resistance Transfer in Breast Cancer. J. Surg. Oncol. 2017, 115, 932–940. [Google Scholar] [CrossRef]
  373. Kreger, B.T.; Johansen, E.R.; Cerione, R.A.; Antonyak, M.A. The Enrichment of Survivin in Exosomes from Breast Cancer Cells Treated with Paclitaxel Promotes Cell Survival and Chemoresistance. Cancers 2016, 8, 111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  374. Chen, W.; Liu, X.; Lv, M.; Chen, L.; Zhao, J.; Zhong, S.; Ji, M.; Hu, Q.; Luo, Z.; Wu, J.; et al. Exosomes from Drug-Resistant Breast Cancer Cells Transmit Chemoresistance by a Horizontal Transfer of MicroRNAs. PLoS ONE 2014, 9, e95240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  375. Santos, J.C.; Lima, N.d.S.; Sarian, L.O.; Matheu, A.; Ribeiro, M.L.; Derchain, S.F.M. Exosome-Mediated Breast Cancer Chemoresistance via MiR-155 Transfer. Sci. Rep. 2018, 8, 829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  376. Yang, L.; Wu, X.; Liang, Y.; Ye, G.; Che, Y.; Wu, X.; Zhu, X.; Fan, H.; Fan, X.; Xu, J. MiR-155 Increases Stemness and Decitabine Resistance in Triple-negative Breast Cancer Cells by Inhibiting TSPAN5. Mol. Carcinog. 2020, 59, 447–461. [Google Scholar] [CrossRef]
  377. Li, X.J.; Ren, Z.J.; Tang, J.H.; Yu, Q. Exosomal MicroRNA MiR-1246 Promotes Cell Proliferation, Invasion and Drug Resistance by Targeting CCNG2 in Breast Cancer. Cell Physiol. Biochem. 2018, 44, 1741–1748. [Google Scholar] [CrossRef]
  378. Wang, B.; Wang, Y.; Wang, X.; Gu, J.; Wu, W.; Wu, H.; Wang, Q.; Zhou, D. Extracellular Vesicles Carrying MiR-887-3p Promote Breast Cancer Cell Drug Resistance by Targeting BTBD7 and Activating the Notch1/Hes1 Signaling Pathway. Dis. Markers 2022, 2022, 5762686. [Google Scholar] [CrossRef] [PubMed]
  379. Wang, B.; Zhang, Y.; Ye, M.; Wu, J.; Ma, L.; Chen, H. Cisplatin-Resistant MDA-MB-231 Cell-Derived Exosomes Increase the Resistance of Recipient Cells in an Exosomal MiR-423-5p-Dependent Manner. Curr. Drug Metab. 2019, 20, 804–814. [Google Scholar] [CrossRef] [PubMed]
  380. Wang, X.; Pei, X.; Guo, G.; Qian, X.; Dou, D.; Zhang, Z.; Xu, X.; Duan, X. Exosome-mediated Transfer of Long Noncoding RNA H19 Induces Doxorubicin Resistance in Breast Cancer. J. Cell Physiol. 2020, 235, 6896–6904. [Google Scholar] [CrossRef]
  381. Zhong, S.; Chen, X.; Wang, D.; Zhang, X.; Shen, H.; Yang, S.; Lv, M.; Tang, J.; Zhao, J. MicroRNA Expression Profiles of Drug-Resistance Breast Cancer Cells and Their Exosomes. Oncotarget 2016, 7, 19601–19609. [Google Scholar] [CrossRef] [Green Version]
  382. Kavanagh, E.L.; Lindsay, S.; Halasz, M.; Gubbins, L.C.; Weiner-Gorzel, K.; Guang, M.H.Z.; McGoldrick, A.; Collins, E.; Henry, M.; Blanco-Fernández, A.; et al. Protein and Chemotherapy Profiling of Extracellular Vesicles Harvested from Therapeutic Induced Senescent Triple Negative Breast Cancer Cells. Oncogenesis 2017, 6, e388. [Google Scholar] [CrossRef] [Green Version]
  383. Chen, W.; Xu, L.; Qian, Q.; He, X.; Peng, W.; Zhu, Y.; Cheng, L. Analysis of MiRNA Signature Differentially Expressed in Exosomes from Adriamycin-Resistant and Parental Human Breast Cancer Cells. Biosci. Rep. 2018, 38, BSR20181090. [Google Scholar] [CrossRef] [Green Version]
  384. Chen, W.-X.; Xu, L.-Y.; Cheng, L.; Qian, Q.; He, X.; Peng, W.-T.; Zhu, Y.-L. Bioinformatics Analysis of Dysregulated MicroRNAs in Exosomes from Docetaxel-Resistant and Parental Human Breast Cancer Cells. Cancer Manag. Res. 2019, 11, 5425–5435. [Google Scholar] [CrossRef] [Green Version]
  385. Koh, M.Z.; Ho, W.Y.; Yeap, S.K.; Ali, N.M.; Yong, C.Y.; Boo, L.; Alitheen, N.B. Exosomal-MicroRNA Transcriptome Profiling of Parental and CSC-like MDA-MB-231 Cells in Response to Cisplatin Treatment. Pathol.-Res. Pract. 2022, 233, 153854. [Google Scholar] [CrossRef] [PubMed]
  386. He, D.; Zhao, Z.; Fu, B.; Li, X.; Zhao, L.; Chen, Y.; Liu, L.; Liu, R.; Li, J. Exosomes Participate in the Radiotherapy Resistance of Cancers. Radiat. Res. 2022, 197, 559–565. [Google Scholar] [CrossRef]
  387. Payton, C.; Pang, L.Y.; Gray, M.; Argyle, D.J. Exosomes Derived from Radioresistant Breast Cancer Cells Promote Therapeutic Resistance in Naïve Recipient Cells. J. Pers. Med. 2021, 11, 1310. [Google Scholar] [CrossRef]
  388. Mutschelknaus, L.; Azimzadeh, O.; Heider, T.; Winkler, K.; Vetter, M.; Kell, R.; Tapio, S.; Merl-Pham, J.; Huber, S.M.; Edalat, L.; et al. Radiation Alters the Cargo of Exosomes Released from Squamous Head and Neck Cancer Cells to Promote Migration of Recipient Cells. Sci. Rep. 2017, 7, 12423. [Google Scholar] [CrossRef] [Green Version]
  389. Drucker, A.; Yoo, B.H.; Khan, I.A.; Choi, D.; Montermini, L.; Liu, X.; Jovanovic, S.; Younis, T.; Rosen, K.V. Trastuzumab-Induced Upregulation of a Protein Set in Extracellular Vesicles Emitted by ErbB2-Positive Breast Cancer Cells Correlates with Their Trastuzumab Sensitivity. Breast Cancer Res. 2020, 22, 105. [Google Scholar] [CrossRef]
  390. Jabbari, N.; Nawaz, M.; Rezaie, J. Ionizing Radiation Increases the Activity of Exosomal Secretory Pathway in MCF-7 Human Breast Cancer Cells: A Possible Way to Communicate Resistance against Radiotherapy. Int. J. Mol. Sci. 2019, 20, 3649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  391. Diamond, J.M.; Vanpouille-Box, C.; Spada, S.; Rudqvist, N.-P.; Chapman, J.R.; Ueberheide, B.M.; Pilones, K.A.; Sarfraz, Y.; Formenti, S.C.; Demaria, S. Exosomes Shuttle TREX1-Sensitive IFN-Stimulatory DsDNA from Irradiated Cancer Cells to DCs. Cancer Immunol. Res. 2018, 6, 910–920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  392. Clark, G.C.; Hampton, J.D.; Koblinski, J.E.; Quinn, B.; Mahmoodi, S.; Metcalf, O.; Guo, C.; Peterson, E.; Fisher, P.B.; Farrell, N.P.; et al. Radiation Induces ESCRT Pathway Dependent CD44v3+ Extracellular Vesicle Production Stimulating Pro-Tumor Fibroblast Activity in Breast Cancer. Front. Oncol. 2022, 12, 913656. [Google Scholar] [CrossRef] [PubMed]
  393. Cortesi, L.; Toss, A.; Cirilli, C.; Marcheselli, L.; Braghiroli, B.; Sebastiani, F.; Federico, M. Twenty-years Experience with de Novo Metastatic Breast Cancer. Int. J. Cancer 2015, 137, 1417–1426. [Google Scholar] [CrossRef]
  394. Zhang, K.; Liu, P.; Tang, H.; Xie, X.; Kong, Y.; Song, C.; Qiu, X.; Xiao, X. AFAP1-AS1 Promotes Epithelial-Mesenchymal Transition and Tumorigenesis Through Wnt/β-Catenin Signaling Pathway in Triple-Negative Breast Cancer. Front. Pharmacol. 2018, 9, 1248. [Google Scholar] [CrossRef] [Green Version]
  395. Zhang, X.; Li, F.; Zhou, Y.; Mao, F.; Lin, Y.; Shen, S.; Li, Y.; Zhang, S.; Sun, Q. Long noncoding RNA AFAP1-AS1 promotes tumor progression and invasion by regulating the miR-2110/Sp1 axis in triple-negative breast cancer. Cell Death Dis. 2020, 12, 627. [Google Scholar] [CrossRef] [PubMed]
  396. Zheng, Z.; Chen, M.; Xing, P.; Yan, X.; Xie, B. Increased Expression of Exosomal AGAP2-AS1 (AGAP2 Antisense RNA 1) In Breast Cancer Cells Inhibits Trastuzumab-Induced Cell Cytotoxicity. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2019, 25, 2211–2220. [Google Scholar] [CrossRef] [PubMed]
  397. Zhang, H.; Yan, C.; Wang, Y. Exosome-Mediated Transfer of CircHIPK3 Promotes Trastuzumab Chemoresistance in Breast Cancer. J. Drug Target. 2021, 29, 1004–1015. [Google Scholar] [CrossRef] [PubMed]
  398. Zhang, Z.; Zhang, L.; Yu, G.; Sun, Z.; Wang, T.; Tian, X.; Duan, X.; Zhang, C. Exosomal MiR-1246 and MiR-155 as Predictive and Prognostic Biomarkers for Trastuzumab-Based Therapy Resistance in HER2-Positive Breast Cancer. Cancer Chemother. Pharmacol. 2020, 86, 761–772. [Google Scholar] [CrossRef]
  399. Martinez, V.G.; O’Neill, S.; Salimu, J.; Breslin, S.; Clayton, A.; Crown, J.; O’Driscoll, L. Resistance to HER2-Targeted Anti-Cancer Drugs Is Associated with Immune Evasion in Cancer Cells and Their Derived Extracellular Vesicles. Oncoimmunology 2017, 6, e1362530. [Google Scholar] [CrossRef]
  400. Cornell, L.; Wander, S.A.; Visal, T.; Wagle, N.; Shapiro, G.I. MicroRNA-Mediated Suppression of the TGF-β Pathway Confers Transmissible and Reversible CDK4/6 Inhibitor Resistance. Cell Rep. 2019, 26, 2667–2680.e7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  401. Re, M.D.; Bertolini, I.; Crucitta, S.; Fontanelli, L.; Rofi, E.; Angelis, C.D.; Diodati, L.; Cavallero, D.; Gianfilippo, G.; Salvadori, B.; et al. Overexpression of TK1 and CDK9 in Plasma-Derived Exosomes Is Associated with Clinical Resistance to CDK4/6 Inhibitors in Metastatic Breast Cancer Patients. Breast Cancer Res. Treat. 2019, 178, 57–62. [Google Scholar] [CrossRef]
  402. Zhang, Z.; Ma, X.; Ekiert, J.; Simons, Y.; Coleman, L.; Truica, C.I.; Blaes, A.H.; Rana, J.; Tandra, P.; Green, L.; et al. Identification of Exosome Protein Biomarkers in Patients with Advanced Hormone Receptor-Positive Breast Cancer Treated with Palbociclib and Tamoxifen. J. Clin. Oncol. 2022, 40, e13014. [Google Scholar] [CrossRef]
  403. Parise, C.A.; Caggiano, V. Risk of Mortality of Node-Negative, ER/PR/HER2 Breast Cancer Subtypes in T1, T2, and T3 Tumors. Breast Cancer Res. Treat. 2017, 165, 743–750. [Google Scholar] [CrossRef]
  404. Tirada, N.; Aujero, M.; Khorjekar, G.; Richards, S.; Chopra, J.; Dromi, S.; Ioffe, O. Breast Cancer Tissue Markers, Genomic Profiling, and Other Prognostic Factors: A Primer for Radiologists. Radiographics 2018, 38, 1902–1920. [Google Scholar] [CrossRef] [Green Version]
  405. Seale, K.N.; Tkaczuk, K.H.R. Circulating Biomarkers in Breast Cancer. Clin. Breast Cancer 2021, 22, e319–e331. [Google Scholar] [CrossRef]
  406. Veyssière, H.; Bidet, Y.; Penault-Llorca, F.; Radosevic-Robin, N.; Durando, X. Circulating Proteins as Predictive and Prognostic Biomarkers in Breast Cancer. Clin. Proteom. 2022, 19, 25. [Google Scholar] [CrossRef] [PubMed]
  407. Tay, T.K.Y.; Tan, P.H. Liquid Biopsy in Breast Cancer: A Focused Review. Arch. Pathol. Lab. Med. 2020, 145, 678–686. [Google Scholar] [CrossRef] [PubMed]
  408. Nassar, F.J.; Chamandi, G.; Tfaily, M.A.; Zgheib, N.K.; Nasr, R. Peripheral Blood-Based Biopsy for Breast Cancer Risk Prediction and Early Detection. Front. Med. 2020, 7, 28. [Google Scholar] [CrossRef]
  409. Jin, Y.; Chen, K.; Wang, Z.; Wang, Y.; Liu, J.; Lin, L.; Shao, Y.; Gao, L.; Yin, H.; Cui, C.; et al. DNA in Serum Extracellular Vesicles Is Stable under Different Storage Conditions. BMC Cancer 2016, 16, 753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  410. Shao, H.; Im, H.; Castro, C.M.; Breakefield, X.; Weissleder, R.; Lee, H. New Technologies for Analysis of Extracellular Vesicles. Chem. Rev. 2018, 118, 1917–1950. [Google Scholar] [CrossRef]
  411. Ayala-Mar, S.; Donoso-Quezada, J.; Gallo-Villanueva, R.C.; Perez-Gonzalez, V.H.; González-Valdez, J. Recent Advances and Challenges in the Recovery and Purification of Cellular Exosomes. Electrophoresis 2019, 40, 3036–3049. [Google Scholar] [CrossRef] [Green Version]
  412. Sidhom, K.; Obi, P.O.; Saleem, A. A Review of Exosomal Isolation Methods: Is Size Exclusion Chromatography the Best Option? Int. J. Mol. Sci. 2020, 21, 6466. [Google Scholar] [CrossRef]
  413. Chen, J.; Li, P.; Zhang, T.; Xu, Z.; Huang, X.; Wang, R.; Du, L. Review on Strategies and Technologies for Exosome Isolation and Purification. Front. Bioeng. Biotechnol. 2022, 9, 811971. [Google Scholar] [CrossRef]
  414. Kumar, D.N.; Chaudhuri, A.; Aqil, F.; Dehari, D.; Munagala, R.; Singh, S.; Gupta, R.C.; Agrawal, A.K. Exosomes as Emerging Drug Delivery and Diagnostic Modality for Breast Cancer: Recent Advances in Isolation and Application. Cancers 2022, 14, 1435. [Google Scholar] [CrossRef] [PubMed]
  415. Valadi, H.; Ekström, K.; Bossios, A.; Sjöstrand, M.; Lee, J.J.; Lötvall, J.O. Exosome-Mediated Transfer of MRNAs and MicroRNAs Is a Novel Mechanism of Genetic Exchange between Cells. Nat. Cell Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  416. Prieto-Vila, M.; Yoshioka, Y.; Ochiya, T. Biological Functions Driven by MRNAs Carried by Extracellular Vesicles in Cancer. Front. Cell Dev. Biol. 2021, 9, 620498. [Google Scholar] [CrossRef] [PubMed]
  417. Zomer, A.; Maynard, C.; Verweij, F.J.; Kamermans, A.; Schäfer, R.; Beerling, E.; Schiffelers, R.M.; de Wit, E.; Berenguer, J.; Ellenbroek, S.I.J.; et al. In Vivo Imaging Reveals Extracellular Vesicle-Mediated Phenocopying of Metastatic Behavior. Cell 2015, 161, 1046–1057. [Google Scholar] [CrossRef] [Green Version]
  418. Kim, K.M.; Abdelmohsen, K.; Mustapic, M.; Kapogiannis, D.; Gorospe, M. RNA in Extracellular Vesicles: RNA in Extracellular Vesicles. Wiley Interdiscip. Rev. RNA 2017, 8, e1413. [Google Scholar] [CrossRef]
  419. Jenjaroenpun, P.; Kremenska, Y.; Nair, V.M.; Kremenskoy, M.; Joseph, B.; Kurochkin, I.V. Characterization of RNA in Exosomes Secreted by Human Breast Cancer Cell Lines Using Next-Generation Sequencing. PeerJ 2013, 1, e201. [Google Scholar] [CrossRef] [Green Version]
  420. Conley, A.; Minciacchi, V.R.; Lee, D.H.; Knudsen, B.S.; Karlan, B.Y.; Citrigno, L.; Viglietto, G.; Tewari, M.; Freeman, M.R.; Demichelis, F.; et al. High-Throughput Sequencing of Two Populations of Extracellular Vesicles Provides an MRNA Signature That Can Be Detected in the Circulation of Breast Cancer Patients. RNA Biol. 2016, 14, 305–316. [Google Scholar] [CrossRef]
  421. 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]
  422. Zhao, F.; Cheng, L.; Shao, Q.; Chen, Z.; Lv, X.; Li, J.; He, L.; Sun, Y.; Ji, Q.; Lu, P.; et al. Characterization of Serum Small Extracellular Vesicles and Their Small RNA Contents across Humans, Rats, and Mice. Sci. Rep. 2020, 10, 4197. [Google Scholar] [CrossRef] [Green Version]
  423. Ozawa, P.M.M.; Jucoski, T.S.; Vieira, E.; Carvalho, T.M.; Malheiros, D.; Ribeiro, E.M.d.S.F. Liquid Biopsy for Breast Cancer Using Extracellular Vesicles and Cell-Free MicroRNAs as Biomarkers. Transl. Res. 2020, 223, 40–60. [Google Scholar] [CrossRef]
  424. Sempere, L.F.; Keto, J.; Fabbri, M. Exosomal MicroRNAs in Breast Cancer towards Diagnostic and Therapeutic Applications. Cancers 2017, 9, 71. [Google Scholar] [CrossRef] [Green Version]
  425. Aggarwal, T.; Wadhwa, R.; Gupta, R.; Paudel, K.R.; Collet, T.; Chellappan, D.K.; Gupta, G.; Perumalsamy, H.; Mehta, M.; Satija, S.; et al. MicroRNAs as Biomarker for Breast Cancer. Endocr. Metab. Immune Disord.-Drug Targets 2019, 20, 1597–1610. [Google Scholar] [CrossRef]
  426. Khalife, H.; Skafi, N.; Fayyad-Kazan, M.; Badran, B. MicroRNAs in Breast Cancer: New Maestros Defining the Melody. Cancer Genet. 2020, 246–247, 18–40. [Google Scholar] [CrossRef] [PubMed]
  427. Richard, V.; Davey, M.G.; Annuk, H.; Miller, N.; Dwyer, R.M.; Lowery, A.; Kerin, M.J. MicroRNAs in Molecular Classification and Pathogenesis of Breast Tumors. Cancers 2021, 13, 5332. [Google Scholar] [CrossRef] [PubMed]
  428. Shi, J. Considering Exosomal MiR-21 as a Biomarker for Cancer. J. Clin. Med. 2016, 5, 42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  429. Hou, Y.; Wang, J.; Wang, X.; Shi, S.; Wang, W.; Chen, Z. Appraising MicroRNA-155 as a Noninvasive Diagnostic Biomarker for Cancer Detection. Medicine 2016, 95, e2450. [Google Scholar] [CrossRef] [PubMed]
  430. Yoshikawa, M.; Iinuma, H.; Umemoto, Y.; Yanagisawa, T.; Matsumoto, A.; Jinno, H. Exosome-Encapsulated MicroRNA-223-3p as a Minimally Invasive Biomarker for the Early Detection of Invasive Breast Cancer. Oncol. Lett. 2018, 15, 9584–9592. [Google Scholar] [CrossRef] [Green Version]
  431. Eichelser, C.; Stückrath, I.; Müller, V.; Milde-Langosch, K.; Wikman, H.; Pantel, K.; Schwarzenbach, H. Increased Serum Levels of Circulating Exosomal MicroRNA-373 in Receptor-Negative Breast Cancer Patients. Oncotarget 2014, 5, 9650–9663. [Google Scholar] [CrossRef] [Green Version]
  432. Shen, S.; Song, Y.; Zhao, B.; Xu, Y.; Ren, X.; Zhou, Y.; Sun, Q. Cancer-Derived Exosomal MiR-7641 Promotes Breast Cancer Progression and Metastasis. Cell Commun. Signal. 2021, 19, 20. [Google Scholar] [CrossRef]
  433. Hannafon, B.N.; Trigoso, Y.D.; Calloway, C.L.; Zhao, Y.D.; Lum, D.H.; Welm, A.L.; Zhao, Z.J.; Blick, K.E.; Dooley, W.C.; Ding, W.Q. Plasma Exosome MicroRNAs Are Indicative of Breast Cancer. Breast Cancer Res. 2016, 18, 90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  434. Motawi, T.M.K.; Sadik, N.A.H.; Shaker, O.G.; Masry, M.R.E.; Mohareb, F. Study of MicroRNAs-21/221 as Potential Breast Cancer Biomarkers in Egyptian Women. Gene 2016, 590, 210–219. [Google Scholar] [CrossRef] [Green Version]
  435. Ali, S.A.; Abdulrahman, Z.F.A.; Faraidun, H.N. Circulatory MiRNA-155, MiRNA-21 Target PTEN Expression and Activity as a Factor in Breast Cancer Development. Cell Mol. Biol. 2020, 66, 44–50. [Google Scholar] [CrossRef]
  436. Konoshenko, M.; Sagaradze, G.; Orlova, E.; Shtam, T.; Proskura, K.; Kamyshinsky, R.; Yunusova, N.; Alexandrova, A.; Efimenko, A.; Tamkovich, S. Total Blood Exosomes in Breast Cancer: Potential Role in Crucial Steps of Tumorigenesis. Int. J. Mol. Sci. 2020, 21, 7341. [Google Scholar] [CrossRef] [PubMed]
  437. Ni, Q.; Stevic, I.; Pan, C.; Müller, V.; Oliveira-Ferrer, L.; Pantel, K.; Schwarzenbach, H. Different Signatures of MiR-16, MiR-30b and MiR-93 in Exosomes from Breast Cancer and DCIS Patients. Sci. Rep. 2018, 8, 12974. [Google Scholar] [CrossRef] [PubMed]
  438. Rodríguez-Martínez, A.; de Miguel-Pérez, D.; Ortega, F.G.; García-Puche, J.L.; Robles-Fernández, I.; Exposito, J.; Martorell-Marugan, J.; Carmona-Sáez, P.; Garrido-Navas, M.d.C.; Rolfo, C.; et al. Exosomal MiRNA Profile as Complementary Tool in the Diagnostic and Prediction of Treatment Response in Localized Breast Cancer under Neoadjuvant Chemotherapy. Breast Cancer Res. 2019, 21, 21. [Google Scholar] [CrossRef] [PubMed]
  439. Han, J.-G.; Jiang, Y.-D.; Zhang, C.-H.; Yang, Y.-M.; Pang, D.; Song, Y.-N.; Zhang, G.-Q. A Novel Panel of Serum MiR-21/MiR-155/MiR-365 as a Potential Diagnostic Biomarker for Breast Cancer. Ann. Surg. Treat. Res. 2017, 92, 55–66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  440. Gonzalez-Villasana, V.; Rashed, M.H.; Gonzalez-Cantú, Y.; Bayraktar, R.; Menchaca-Arredondo, J.L.; Vazquez-Guillen, J.M.; Rodriguez-Padilla, C.; Lopez-Berestein, G.; Resendez-Perez, D. Presence of Circulating MiR-145, MiR-155, and MiR-382 in Exosomes Isolated from Serum of Breast Cancer Patients and Healthy Donors. Dis. Markers 2019, 2019, 6852917. [Google Scholar] [CrossRef] [PubMed]
  441. Khalighfard, S.; Alizadeh, A.M.; Irani, S.; Omranipour, R. Plasma MiR-21, MiR-155, MiR-10b, and Let-7a as the Potential Biomarkers for the Monitoring of Breast Cancer Patients. Sci. Rep. 2018, 8, 17981. [Google Scholar] [CrossRef] [Green Version]
  442. Fang, R.; Zhu, Y.; Hu, L.; Khadka, V.S.; Ai, J.; Zou, H.; Ju, D.; Jiang, B.; Deng, Y.; Hu, X. Plasma MicroRNA Pair Panels as Novel Biomarkers for Detection of Early Stage Breast Cancer. Front. Physiol. 2019, 9, 1879. [Google Scholar] [CrossRef] [Green Version]
  443. Shamsizadeh, S.; Goliaei, S.; Moghadam, Z.R. CAMIRADA: Cancer MicroRNA Association Discovery Algorithm, a Case Study on Breast Cancer. J. Biomed. Inform. 2019, 94, 103180. [Google Scholar] [CrossRef] [Green Version]
  444. Lv, S.; Wang, Y.; Xu, W.; Dong, X. Serum Exosomal MiR-17-5p as a Promising Biomarker Diagnostic Biomarker for Breast Cancer. Clin. Lab. 2020, 66. [Google Scholar] [CrossRef]
  445. Shimomura, A.; Shiino, S.; Kawauchi, J.; Takizawa, S.; Sakamoto, H.; Matsuzaki, J.; Ono, M.; Takeshita, F.; Niida, S.; Shimizu, C.; et al. Novel Combination of Serum MicroRNA for Detecting Breast Cancer in the Early Stage. Cancer Sci. 2016, 107, 326–334. [Google Scholar] [CrossRef] [Green Version]
  446. Asadirad, A.; Khodadadi, A.; Talaiezadeh, A.; Shohan, M.; Rashno, M.; Joudaki, N. Evaluation of MiRNA-21-5p and MiRNA-10b-5p Levels in Serum-Derived Exosomes of Breast Cancer Patients in Different Grades. Mol. Cell Probes 2022, 64, 101831. [Google Scholar] [CrossRef] [PubMed]
  447. Hamam, R.; Ali, A.M.; Alsaleh, K.A.; Kassem, M.; Alfayez, M.; Aldahmash, A.; Alajez, N.M. MicroRNA Expression Profiling on Individual Breast Cancer Patients Identifies Novel Panel of Circulating MicroRNA for Early Detection. Sci. Rep. 2016, 6, 25997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  448. Liu, D.; Li, B.; Shi, X.; Zhang, J.; Chen, A.M.; Xu, J.; Wang, W.; Huang, K.; Gao, J.; Zheng, Z.; et al. Cross-Platform Genomic Identification and Clinical Validation of Breast Cancer Diagnostic Biomarkers. Aging 2021, 13, 4258–4273. [Google Scholar] [CrossRef] [PubMed]
  449. Yan, C.; Hu, J.; Yang, Y.; Hu, H.; Zhou, D.; Ma, M.; Xu, N. Plasma Extracellular Vesicle-Packaged MicroRNAs as Candidate Diagnostic Biomarkers for Early-Stage Breast Cancer. Mol. Med. Rep. 2019, 20, 3991–4002. [Google Scholar] [CrossRef] [PubMed]
  450. Adam-Artigues, A.; Garrido-Cano, I.; Carbonell-Asins, J.A.; Lameirinhas, A.; Simón, S.; Ortega-Morillo, B.; Martínez, M.T.; Hernando, C.; Constâncio, V.; Burgues, O.; et al. Identification of a Two-MicroRNA Signature in Plasma as a Novel Biomarker for Very Early Diagnosis of Breast Cancer. Cancers 2021, 13, 2848. [Google Scholar] [CrossRef]
  451. Jang, J.Y.; Kim, Y.S.; Kang, K.N.; Kim, K.H.; Park, Y.J.; Kim, C.W. Multiple MicroRNAs as Biomarkers for Early Breast Cancer Diagnosis. Mol. Clin. Oncol. 2021, 14, 31. [Google Scholar] [CrossRef] [PubMed]
  452. Borsos, B.N.; Páhi, Z.G.; Ujfaludi, Z.; Sükösd, F.; Nikolényi, A.; Bankó, S.; Pankotai-Bodó, G.; Oláh-Németh, O.; Pankotai, T. BC-MiR: Monitoring Breast Cancer-Related MiRNA Profile in Blood Sera—A Prosperous Approach for Tumor Detection. Cells 2022, 11, 2721. [Google Scholar] [CrossRef] [PubMed]
  453. Zografos, E.; Zagouri, F.; Kalapanida, D.; Zakopoulou, R.; Kyriazoglou, A.; Apostolidou, K.; Gazouli, M.; Dimopoulos, M.-A. Prognostic Role of MicroRNAs in Breast Cancer: A Systematic Review. Oncotarget 2019, 10, 7156–7178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  454. Pourteimoor, V.; Paryan, M.; Mohammadi-Yeganeh, S. MicroRNA as a Systemic Intervention in the Specific Breast Cancer Subtypes with C-MYC Impacts; Introducing Subtype-based Appraisal Tool. J. Cell Physiol. 2018, 233, 5655–5669. [Google Scholar] [CrossRef] [PubMed]
  455. Stevic, I.; Müller, V.; Weber, K.; Fasching, P.A.; Karn, T.; Marmé, F.; Schem, C.; Stickeler, E.; Denkert, C.; van Mackelenbergh, M.; et al. Specific MicroRNA Signatures in Exosomes of Triple-Negative and HER2-Positive Breast Cancer Patients Undergoing Neoadjuvant Therapy within the GeparSixto Trial. BMC Med. 2018, 16, 179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  456. Huo, D.; Clayton, W.M.; Yoshimatsu, T.F.; Chen, J.; Olopade, O.I. Identification of a Circulating MicroRNA Signature to Distinguish Recurrence in Breast Cancer Patients. Oncotarget 2016, 7, 55231–55248. [Google Scholar] [CrossRef] [Green Version]
  457. Papadaki, C.; Thomopoulou, K.; Koronakis, G.; Monastirioti, A.A.; Papadaki, M.A.; Kalbakis, K.; Agelaki, S.; Mavroudis, D. MicroRNAs Involved in Immune Response as Prognostic Markers in Early and Metastatic Breast Cancer. J. Clin. Oncol. 2020, 38, e15528. [Google Scholar] [CrossRef]
  458. Thomopoulou, K.; Papadaki, C.; Monastirioti, A.; Koronakis, G.; Mala, A.; Kalapanida, D.; Mavroudis, D.; Agelaki, S. MicroRNAs Regulating Tumor Immune Response in the Prediction of the Outcome in Patients With Breast Cancer. Front. Mol. Biosci. 2021, 8, 668534. [Google Scholar] [CrossRef]
  459. Joshi, S.; Garlapati, C.; Bhattarai, S.; Su, Y.; Rios-Colon, L.; Deep, G.; Torres, M.A.; Aneja, R. Exosomal Metabolic Signatures Are Associated with Differential Response to Neoadjuvant Chemotherapy in Patients with Breast Cancer. Int. J. Mol. Sci. 2022, 23, 5324. [Google Scholar] [CrossRef]
  460. Todorova, V.K.; Byrum, S.D.; Gies, A.J.; Haynie, C.; Smith, H.; Reyna, N.S.; Makhoul, I. Circulating Exosomal MicroRNAs as Predictive Biomarkers of Neoadjuvant Chemotherapy Response in Breast Cancer. Curr. Oncol. 2022, 29, 55. [Google Scholar] [CrossRef]
  461. Zhang, Z.; Zhang, H.; Yu, J.; Xu, L.; Pang, X.; Xiang, Q.; Liu, Q.; Cui, Y. MiRNAs as Therapeutic Predictors and Prognostic Biomarkers of Neoadjuvant Chemotherapy in Breast Cancer: A Systematic Review and Meta-Analysis. Breast Cancer Res. Treat. 2022, 194, 483–505. [Google Scholar] [CrossRef]
  462. Zhu, W.; Liu, M.; Fan, Y.; Ma, F.; Xu, N.; Xu, B. Dynamics of Circulating MicroRNAs as a Novel Indicator of Clinical Response to Neoadjuvant Chemotherapy in Breast Cancer. Cancer Med. 2018, 7, 4420–4433. [Google Scholar] [CrossRef] [Green Version]
  463. Sueta, A.; Fujiki, Y.; Goto-Yamaguchi, L.; Tomiguchi, M.; Yamamoto-Ibusuki, M.; Iwase, H.; Yamamoto, Y. Exosomal MiRNA Profiles of Triple-Negative Breast Cancer in Neoadjuvant Treatment. Oncol. Lett. 2021, 22, 819. [Google Scholar] [CrossRef] [PubMed]
  464. Li, S.; Yang, X.; Yang, J.; Zhen, J.; Zhang, D. Serum MicroRNA-21 as a Potential Diagnostic Biomarker for Breast Cancer: A Systematic Review and Meta-Analysis. Clin. Exp. Med. 2016, 16, 29–35. [Google Scholar] [CrossRef]
  465. Salim, B.; Athira, M.V.; Kandaswamy, A.; Vijayakumar, M. Investigation on Staging of Breast Cancer Using MiR-21 as a Biomarker in the Serum. Int. J. Biomed. Eng. Technol. 2020, 33, 211–221. [Google Scholar] [CrossRef]
  466. Allah, A.W.; Yahya, M.; Elsaeidy, K.S.; Alkanj, S.; Hamam, K.; El-Saady, M.; Ebada, M.A. Clinical Assessment of MiRNA-23b as a Prognostic Factor for Various Carcinomas: A Systematic Review and Meta-Analysis. Meta Gene 2020, 24, 100651. [Google Scholar] [CrossRef]
  467. Li, X.; Zeng, Z.; Wang, J.; Wu, Y.; Chen, W.; Zheng, L.; Xi, T.; Wang, A.; Lu, Y. MicroRNA-9 and Breast Cancer. Biomed. Pharmacother. 2020, 122, 109687. [Google Scholar] [CrossRef] [PubMed]
  468. Sehovic, E.; Urru, S.; Chiorino, G.; Doebler, P. Meta-Analysis of Diagnostic Cell-Free Circulating MicroRNAs for Breast Cancer Detection. BMC Cancer 2022, 22, 634. [Google Scholar] [CrossRef]
  469. Hamam, R.; Hamam, D.; Alsaleh, K.A.; Kassem, M.; Zaher, W.; Alfayez, M.; Aldahmash, A.; Alajez, N.M. Circulating MicroRNAs in Breast Cancer: Novel Diagnostic and Prognostic Biomarkers. Cell Death Dis. 2017, 8, e3045. [Google Scholar] [CrossRef] [Green Version]
  470. Li, J.; Guan, X.; Fan, Z.; Ching, L.-M.; Li, Y.; Wang, X.; Cao, W.-M.; Liu, D.-X. Non-Invasive Biomarkers for Early Detection of Breast Cancer. Cancers 2020, 12, 2767. [Google Scholar] [CrossRef]
  471. Yao, R.-W.; Wang, Y.; Chen, L.-L. Cellular Functions of Long Noncoding RNAs. Nat. Cell Biol. 2019, 21, 542–551. [Google Scholar] [CrossRef]
  472. Qian, Y.; Shi, L.; Luo, Z. Long Non-Coding RNAs in Cancer: Implications for Diagnosis, Prognosis, and Therapy. Front. Med. 2020, 7, 612393. [Google Scholar] [CrossRef]
  473. Liu, S.J.; Dang, H.X.; Lim, D.A.; Feng, F.Y.; Maher, C.A. Long Noncoding RNAs in Cancer Metastasis. Nat. Rev. Cancer 2021, 21, 446–460. [Google Scholar] [CrossRef] [PubMed]
  474. Ming, H.; Li, B.; Zhou, L.; Goel, A.; Huang, C. Long Non-Coding RNAs and Cancer Metastasis: Molecular Basis and Therapeutic Implications. Biochim. Biophys. Acta BBA-Rev. Cancer 2021, 1875, 188519. [Google Scholar] [CrossRef] [PubMed]
  475. Shi, W.; Jin, X.; Wang, Y.; Zhang, Q.; Yang, L. High Serum Exosomal Long Non-coding RNA DANCR Expression Confers Poor Prognosis in Patients with Breast Cancer. J. Clin. Lab. Anal. 2022, 36, e24186. [Google Scholar] [CrossRef] [PubMed]
  476. Zhang, P.; Zhou, H.; Lu, K.; Lu, Y.; Wang, Y.; Feng, T. Exosome-Mediated Delivery of MALAT1 Induces Cell Proliferation in Breast Cancer. Oncotargets Ther. 2018, 11, 291–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  477. Lan, F.; Zhang, X.; Li, H.; Yue, X.; Sun, Q. Serum Exosomal LncRNA XIST Is a Potential Non-invasive Biomarker to Diagnose Recurrence of Triple-negative Breast Cancer. J. Cell Mol. Med. 2021, 25, 7602–7607. [Google Scholar] [CrossRef]
  478. Qiu, P.; Guo, Q.; Lin, J.; Pan, K.; Chen, J.; Ding, M. An Exosome-Related Long Non-Coding RNAs Risk Model Could Predict Survival Outcomes in Patients with Breast Cancer. Sci. Rep. 2022, 12, 22322. [Google Scholar] [CrossRef] [PubMed]
  479. Yu, Q.; Peng, S.; Li, J.; Tan, X. Exosomal-mediated transfer of OIP5-AS1 enhanced cell chemoresistance to trastuzumab in breast cancer via up-regulating HMGB3 by sponging miR-381-3p. Open Med (Wars) 2021, 16, 512–525. [Google Scholar] [CrossRef]
  480. Shi, S.-J.; Wang, L.-J.; Yu, B.; Li, Y.-H.; Jin, Y.; Bai, X.-Z. LncRNA-ATB Promotes Trastuzumab Resistance and Invasion-Metastasis Cascade in Breast Cancer. Oncotarget 2015, 6, 11652–11663. [Google Scholar] [CrossRef] [Green Version]
  481. Qian, X.; Qu, H.; Zhang, F.; Peng, S.; Dou, D.; Yang, Y.; Ding, Y.; Xie, M.; Dong, H.; Liao, Y.; et al. Exosomal Long Noncoding RNA AGAP2-AS1 Regulates Trastuzumab Resistance via Inducing Autophagy in Breast Cancer. Am. J. Cancer Res. 2020, 11, 1962–1981. [Google Scholar]
  482. Dong, H.; Wang, W.; Chen, R.; Zhang, Y.; Zou, K.; Ye, M.; He, X.; Zhang, F.; Han, J. Exosome-Mediated Transfer of LncRNA-SNHG14 Promotes Trastuzumab Chemoresistance in Breast Cancer. Int. J. Oncol. 2022, 61, 92. [Google Scholar] [CrossRef]
  483. Wu, Z.; Xu, Z.; Yu, B.; Zhang, J.; Yu, B. The Potential Diagnostic Value of Exosomal Long Noncoding RNAs in Solid Tumors: A Meta-Analysis and Systematic Review. Biomed. Res. Int. 2020, 2020, 6786875. [Google Scholar] [CrossRef] [PubMed]
  484. Møller, H.D.; Mohiyuddin, M.; Prada-Luengo, I.; Sailani, M.R.; Halling, J.F.; Plomgaard, P.; Maretty, L.; Hansen, A.J.; Snyder, M.P.; Pilegaard, H.; et al. Circular DNA Elements of Chromosomal Origin Are Common in Healthy Human Somatic Tissue. Nat. Commun. 2018, 9, 1069. [Google Scholar] [CrossRef] [Green Version]
  485. HOLDENRIEDER, S.; STIEBER, P. Therapy Control in Oncology by Circulating Nucleosomes. Ann. N. Y. Acad. Sci. 2004, 1022, 211–216. [Google Scholar] [CrossRef]
  486. Li, Z.; Chen, Z.; Hu, G.; Jiang, Y. Roles of Circular RNA in Breast Cancer: Present and Future. Am. J. Transl. Res. 2019, 11, 3945–3954. [Google Scholar]
  487. Suzuki, H.; Tsukahara, T. A View of Pre-MRNA Splicing from RNase R Resistant RNAs. Int. J. Mol. Sci. 2014, 15, 9331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  488. Hansen, T.B.; Jensen, T.I.; Clausen, B.H.; Bramsen, J.B.; Finsen, B.; Damgaard, C.K.; Kjems, J. Natural RNA Circles Function as Efficient MicroRNA Sponges. Nature 2013, 495, 384–388. [Google Scholar] [CrossRef] [PubMed]
  489. Ding, X.; Zheng, J.; Cao, M. Circ_0004771 Accelerates Cell Carcinogenic Phenotypes via Suppressing MiR-1253-Mediated DDAH1 Inhibition in Breast Cancer. Cancer Manag. Res. 2021, 13, 1–11. [Google Scholar] [CrossRef]
  490. Liu, J.; Peng, X.; Liu, Y.; Hao, R.; Zhao, R.; Zhang, L.; Zhao, F.; Liu, Q.; Liu, Y.; Qi, Y. The Diagnostic Value of Serum Exosomal Has_circ_0000615 for Breast Cancer Patients. Int. J. Gen. Med. 2021, 14, 4545–4554. [Google Scholar] [CrossRef]
  491. Turco, C.; Esposito, G.; Iaiza, A.; Goeman, F.; Benedetti, A.; Gallo, E.; Daralioti, T.; Perracchio, L.; Sacconi, A.; Pasanisi, P.; et al. MALAT1-Dependent Hsa_circ_0076611 Regulates Translation Rate in Triple-Negative Breast Cancer. Commun. Biol. 2022, 5, 598. [Google Scholar] [CrossRef]
  492. Yang, S.; Tang, J. Exosomal Circular RNAs Derived from Serum: Promising Biomarkers for Therapeutic Targets and Prognosis of Triple-Negative Breast Cancer (TNBC). J. Clin. Oncol. 2020, 38, 3528. [Google Scholar] [CrossRef]
  493. Yang, S.; Wang, D.; Zhong, S.; Chen, W.; Wang, F.; Zhang, J.; Xu, W.; Xu, D.; Zhang, Q.; Li, J.; et al. Tumor-Derived Exosomal CircPSMA1 Facilitates the Tumorigenesis, Metastasis, and Migration in Triple-Negative Breast Cancer (TNBC) through MiR-637/Akt1/β-Catenin (Cyclin D1) Axis. Cell. Death Dis. 2021, 12, 420. [Google Scholar] [CrossRef]
  494. Halvaei, S.; Daryani, S.; Eslami-S, Z.; Samadi, T.; Jafarbeik-Iravani, N.; Bakhshayesh, T.O.; Majidzadeh-A., K.; Esmaeili, R. Exosomes in Cancer Liquid Biopsy: A Focus on Breast Cancer. Mol. Ther.-Nucleic Acids 2018, 10, 131–141. [Google Scholar] [CrossRef] [Green Version]
  495. Tutanov, O.; Proskura, K.; Kamyshinsky, R.; Shtam, T.; Tsentalovich, Y.; Tamkovich, S. Proteomic Profiling of Plasma and Total Blood Exosomes in Breast Cancer: A Potential Role in Tumor Progression, Diagnosis, and Prognosis. Front. Oncol. 2020, 10, 580891. [Google Scholar] [CrossRef]
  496. Wang, X.; Zhong, W.; Bu, J.; Li, Y.; Li, R.; Nie, R.; Xiao, C.; Ma, K.; Huang, X.; Li, Y. Exosomal Protein CD82 as a Diagnostic Biomarker for Precision Medicine for Breast Cancer. Mol. Carcinog. 2019, 58, 674–685. [Google Scholar] [CrossRef] [PubMed]
  497. Li, S.; Li, X.; Yang, S.; Pi, H.; Li, Z.; Yao, P.; Zhang, Q.; Wang, Q.; Shen, P.; Li, X.; et al. Proteomic Landscape of Exosomes Reveals the Functional Contributions of CD151 in Triple-Negative Breast Cancer. Mol. Cell Proteom. 2021, 20, 100121. [Google Scholar] [CrossRef]
  498. Rupp, A.-K.; Rupp, C.; Keller, S.; Brase, J.C.; Ehehalt, R.; Fogel, M.; Moldenhauer, G.; Marmé, F.; Sültmann, H.; Altevogt, P. Loss of EpCAM Expression in Breast Cancer Derived Serum Exosomes: Role of Proteolytic Cleavage. Gynecol. Oncol. 2011, 122, 437–446. [Google Scholar] [CrossRef]
  499. Galindo-Hernandez, O.; Villegas-Comonfort, S.; Candanedo, F.; González-Vázquez, M.-C.; Chavez-Ocaña, S.; Jimenez-Villanueva, X.; Sierra-Martinez, M.; Salazar, E.P. Elevated Concentration of Microvesicles Isolated from Peripheral Blood in Breast Cancer Patients. Arch. Med. Res. 2013, 44, 208–214. [Google Scholar] [CrossRef]
  500. Jordan, K.R.; Hall, J.K.; Schedin, T.; Borakove, M.; Xian, J.J.; Dzieciatkowska, M.; Lyons, T.R.; Schedin, P.; Hansen, K.C.; Borges, V.F. Extracellular Vesicles from Young Women’s Breast Cancer Patients Drive Increased Invasion of Non-Malignant Cells via the Focal Adhesion Kinase Pathway: A Proteomic Approach. Breast Cancer Res. 2020, 22, 128. [Google Scholar] [CrossRef]
  501. Higginbotham, J.N.; Demory Beckler, M.; Gephart, J.D.; Franklin, J.L.; Bogatcheva, G.; Kremers, G.-J.; Piston, D.W.; Ayers, G.D.; McConnell, R.E.; Tyska, M.J.; et al. Amphiregulin Exosomes Increase Cancer Cell Invasion. Curr. Biol. 2011, 21, 779–786. [Google Scholar] [CrossRef] [Green Version]
  502. Bi, T.-L.; Sun, J.-J.; Tian, Y.-Z.; Zhou, Y.-F. Research Progress of Relationship between Exosomes and Breast Cancer. Sheng Li Xue Bao Acta Physiol. Sin. 2016, 68, 352–358. [Google Scholar]
  503. Moon, P.-G.; Lee, J.-E.; Cho, Y.-E.; Lee, S.J.; Chae, Y.S.; Jung, J.H.; Kim, I.-S.; Park, H.Y.; Baek, M.-C. Fibronectin on Circulating Extracellular Vesicles as a Liquid Biopsy to Detect Breast Cancer. Oncotarget 2016, 7, 40189–40199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  504. Lee, S.J.; Lee, J.; Jung, J.H.; Park, H.Y.; Moon, P.-G.; Chae, Y.S.; Baek, M.-C. Exosomal Del-1 as a Potent Diagnostic Marker for Breast Cancer: Prospective Cohort Study. Clin. Breast Cancer 2021, 21, e748–e756. [Google Scholar] [CrossRef]
  505. Li, K.; Liu, T.; Chen, J.; Ni, H.; Li, W. Survivin in Breast Cancer–Derived Exosomes Activates Fibroblasts by up-Regulating SOD1, Whose Feedback Promotes Cancer Proliferation and Metastasis. J. Biol. Chem. 2020, 295, 13737–13752. [Google Scholar] [CrossRef] [PubMed]
  506. Wang, X.; Cheng, K.; Zhang, G.; Jia, Z.; Yu, Y.; Guo, J.; Hua, Y.; Guo, F.; Li, X.; Zou, W.; et al. Enrichment of CD44 in Exosomes From Breast Cancer Cells Treated With Doxorubicin Promotes Chemoresistance. Front. Oncol. 2020, 10, 960. [Google Scholar] [CrossRef]
  507. Cao, Y.; Wang, Y.; Yu, X.; Jiang, X.; Li, G.; Zhao, J. Identification of Programmed Death Ligand-1 Positive Exosomes in Breast Cancer Based on DNA Amplification-Responsive Metal-Organic Frameworks. Biosens. Bioelectron. 2020, 166, 112452. [Google Scholar] [CrossRef] [PubMed]
  508. Jung, H.H.; Kim, J.-Y.; Cho, E.Y.; Oh, J.M.; Lee, J.E.; Kim, S.W.; Nam, S.J.; Park, Y.H.; Ahn, J.S.; Im, Y.-H. Elevated Level of Nerve Growth Factor (NGF) in Serum-Derived Exosomes Predicts Poor Survival in Patients with Breast Cancer Undergoing Neoadjuvant Chemotherapy. Cancers 2021, 13, 5260. [Google Scholar] [CrossRef] [PubMed]
  509. Bastos, N.; Ruivo, C.F.; da Silva, S.; Melo, S.A. Exosomes in Cancer: Use Them or Target Them? Semin. Cell Dev. Biol. 2018, 78, 13–21. [Google Scholar] [CrossRef]
  510. Mulcahy, L.A.; Pink, R.C.; Carter, D.R.F. Routes and Mechanisms of Extracellular Vesicle Uptake. J. Extracell. Vesicles 2014, 3, 24641. [Google Scholar] [CrossRef] [Green Version]
  511. O’Donoghue, E.J.; Krachler, A.M. Mechanisms of Outer Membrane Vesicle Entry into Host Cells. Cell Microbiol. 2016, 18, 1508–1517. [Google Scholar] [CrossRef]
  512. Christianson, H.C.; Svensson, K.J.; van Kuppevelt, T.H.; Li, J.-P.; Belting, M. Cancer Cell Exosomes Depend on Cell-Surface Heparan Sulfate Proteoglycans for Their Internalization and Functional Activity. Proc. Natl. Acad. Sci. USA 2013, 110, 17380–17385. [Google Scholar] [CrossRef] [Green Version]
  513. Escrevente, C.; Keller, S.; Altevogt, P.; Costa, J. Interaction and uptake of exosomes by ovarian cancer cells. BMC Cancer 2011, 11, 108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  514. Svensson, K.J.; Christianson, H.C.; Wittrup, A.; Bourseau-Guilmain, E.; Lindqvist, E.; Svensson, L.M.; Mörgelin, M.; Belting, M. Exosome Uptake Depends on ERK1/2-Heat Shock Protein 27 Signaling and Lipid Raft-Mediated Endocytosis Negatively Regulated by Caveolin-1. J. Biol. Chem. 2013, 288, 17713–17724. [Google Scholar] [CrossRef] [Green Version]
  515. Wang, L.H.; Rothberg, K.G.; Anderson, R.G. Mis-Assembly of Clathrin Lattices on Endosomes Reveals a Regulatory Switch for Coated Pit Formation. J. Cell Biol. 1993, 123, 1107–1117. [Google Scholar] [CrossRef]
  516. Ehrlich, M.; Boll, W.; van Oijen, A.; Hariharan, R.; Chandran, K.; Nibert, M.L.; Kirchhausen, T. Endocytosis by Random Initiation and Stabilization of Clathrin-Coated Pits. Cell 2004, 118, 591–605. [Google Scholar] [CrossRef] [Green Version]
  517. Koivusalo, M.; Welch, C.; Hayashi, H.; Scott, C.C.; Kim, M.; Alexander, T.; Touret, N.; Hahn, K.M.; Grinstein, S. Amiloride Inhibits Macropinocytosis by Lowering Submembranous PH and Preventing Rac1 and Cdc42 Signaling. J. Cell Biol. 2010, 188, 547–563. [Google Scholar] [CrossRef] [Green Version]
  518. Parolini, I.; Federici, C.; Raggi, C.; Lugini, L.; Palleschi, S.; Milito, A.D.; Coscia, C.; Iessi, E.; Logozzi, M.; Molinari, A.; et al. Microenvironmental PH Is a Key Factor for Exosome Traffic in Tumor Cells. J. Biol. Chem. 2009, 284, 34211–34222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  519. Hait, N.C.; Maiti, A. The Role of Sphingosine-1-Phosphate and Ceramide-1-Phosphate in Inflammation and Cancer. Mediat. Inflamm. 2017, 2017, 4806541. [Google Scholar] [CrossRef] [Green Version]
  520. Trajkovic, K.; Hsu, C.; Chiantia, S.; Rajendran, L.; Wenzel, D.; Wieland, F.; Schwille, P.; Brügger, B.; Simons, M. Ceramide Triggers Budding of Exosome Vesicles into Multivesicular Endosomes. Science 2008, 319, 1244–1247. [Google Scholar] [CrossRef] [PubMed]
  521. Menck, K.; Sönmezer, C.; Worst, T.S.; Schulz, M.; Dihazi, G.H.; Streit, F.; Erdmann, G.; Kling, S.; Boutros, M.; Binder, C.; et al. Neutral Sphingomyelinases Control Extracellular Vesicles Budding from the Plasma Membrane. J. Extracell. Vesicles 2017, 6, 1378056. [Google Scholar] [CrossRef] [PubMed]
  522. Catalano, M.; O’Driscoll, L. Inhibiting Extracellular Vesicles Formation and Release: A Review of EV Inhibitors. J. Extracell. Vesicles 2019, 9, 1703244. [Google Scholar] [CrossRef] [Green Version]
  523. Kosaka, N.; Iguchi, H.; Yoshioka, Y.; Takeshita, F.; Matsuki, Y.; Ochiya, T. Secretory Mechanisms and Intercellular Transfer of MicroRNAs in Living Cells. J. Biol. Chem. 2010, 285, 17442–17452. [Google Scholar] [CrossRef] [Green Version]
  524. Wang, J.; Yeung, B.Z.; Wientjes, M.G.; Cui, M.; Peer, C.J.; Lu, Z.; Figg, W.D.; Woo, S.; Au, J.L.-S. A Quantitative Pharmacology Model of Exosome-Mediated Drug Efflux and Perturbation-Induced Synergy. Pharmaceutics 2021, 13, 997. [Google Scholar] [CrossRef]
  525. Bianco, F.; Perrotta, C.; Novellino, L.; Francolini, M.; Riganti, L.; Menna, E.; Saglietti, L.; Schuchman, E.H.; Furlan, R.; Clementi, E.; et al. Acid Sphingomyelinase Activity Triggers Microparticle Release from Glial Cells. EMBO J. 2009, 28, 1043–1054. [Google Scholar] [CrossRef] [Green Version]
  526. Kosgodage, U.; Trindade, R.; Thompson, P.; Inal, J.; Lange, S. Chloramidine/Bisindolylmaleimide-I-Mediated Inhibition of Exosome and Microvesicle Release and Enhanced Efficacy of Cancer Chemotherapy. Int. J. Mol. Sci. 2017, 18, 1007. [Google Scholar] [CrossRef] [Green Version]
  527. Baietti, M.F.; Zhang, Z.; Mortier, E.; Melchior, A.; Degeest, G.; Geeraerts, A.; Ivarsson, Y.; Depoortere, F.; Coomans, C.; Vermeiren, E.; et al. Syndecan–Syntenin–ALIX Regulates the Biogenesis of Exosomes. Nat. Cell Biol. 2012, 14, 677–685. [Google Scholar] [CrossRef] [PubMed]
  528. Zhang, Y.; Song, M.; Cui, Z.S.; Li, C.Y.; Xue, X.X.; Yu, M.; Lu, Y.; Zhang, S.Y.; Wang, E.H.; Wen, Y.Y. Down-Regulation of TSG101 by Small Interfering RNA Inhibits the Proliferation of Breast Cancer Cells through the MAPK/ERK Signal Pathway. Histol. Histopathol. 2010, 26, 87–94. [Google Scholar] [CrossRef]
  529. Li, Z.; Fang, R.; Fang, J.; He, S.; Liu, T. Functional Implications of Rab27 GTPases in Cancer. Cell Commun. Signal. 2018, 16, 44. [Google Scholar] [CrossRef] [Green Version]
  530. Li, W.; Hu, Y.; Jiang, T.; Han, Y.; Han, G.; Chen, J.; Li, X. Rab27A Regulates Exosome Secretion from Lung Adenocarcinoma Cells A549: Involvement of EPI64. Apmis 2014, 122, 1080–1087. [Google Scholar] [CrossRef] [PubMed]
  531. Roma-Rodrigues, C.; Pereira, F.; de Matos, A.P.A.; Fernandes, M.; Baptista, P.V.; Fernandes, A.R. Smuggling Gold Nanoparticles across Cell Types—A New Role for Exosomes in Gene Silencing. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 1389–1398. [Google Scholar] [CrossRef]
  532. Hoshino, D.; Kirkbride, K.C.; Costello, K.; Clark, E.S.; Sinha, S.; Grega-Larson, N.; Tyska, M.J.; Weaver, A.M. Exosome Secretion Is Enhanced by Invadopodia and Drives Invasive Behavior. Cell Rep. 2013, 5, 1159–1168. [Google Scholar] [CrossRef] [Green Version]
  533. Sinha, S.; Hoshino, D.; Hong, N.H.; Kirkbride, K.C.; Grega-Larson, N.E.; Seiki, M.; Tyska, M.J.; Weaver, A.M. Cortactin Promotes Exosome Secretion by Controlling Branched Actin Dynamics. J. Cell Biol. 2016, 214, 197–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  534. Twafra, S.; Sokolik, C.G.; Sneh, T.; Srikanth, K.D.; Meirson, T.; Genna, A.; Chill, J.H.; Gil-Henn, H. A Novel Pyk2-Derived Peptide Inhibits Invadopodia-Mediated Breast Cancer Metastasis. Oncogene 2023, 42, 278–292. [Google Scholar] [CrossRef]
  535. Huang, M.-B.; Brena, D.; Wu, J.Y.; Roth, W.W.; Owusu, S.; Bond, V.C. Novel Secretion Modification Region (SMR) Peptide Exhibits Anti-Metastatic Properties in Human Breast Cancer Cells. Sci. Rep. 2022, 12, 13204. [Google Scholar] [CrossRef] [PubMed]
  536. Chen, W.-X.; Cheng, L.; Pan, M.; Qian, Q.; Zhu, Y.-L.; Xu, L.-Y.; Ding, Q. D Rhamnose β-Hederin against Human Breast Cancer by Reducing Tumor-Derived Exosomes. Oncol. Lett. 2018, 16, 5172–5178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  537. Wei, Y.; Li, M.; Cui, S.; Wang, D.; Zhang, C.-Y.; Zen, K.; Li, L. Shikonin Inhibits the Proliferation of Human Breast Cancer Cells by Reducing Tumor-Derived Exosomes. Molecules 2016, 21, 777. [Google Scholar] [CrossRef] [PubMed]
  538. Kosgodage, U.S.; Mould, R.; Henley, A.B.; Nunn, A.V.; Guy, G.W.; Thomas, E.L.; Inal, J.M.; Bell, J.D.; Lange, S. Cannabidiol (CBD) Is a Novel Inhibitor for Exosome and Microvesicle (EMV) Release in Cancer. Front. Pharmacol. 2018, 9, 889. [Google Scholar] [CrossRef] [Green Version]
  539. Huang, M.-B.; Gonzalez, R.R.; Lillard, J.; Bond, V.C. Secretion Modification Region-Derived Peptide Blocks Exosome Release and Mediates Cell Cycle Arrest in Breast Cancer Cells. Oncotarget 2017, 8, 11302–11315. [Google Scholar] [CrossRef] [Green Version]
  540. Conlan, R.S.; Pisano, S.; Oliveira, M.I.; Ferrari, M.; Pinto, I.M. Exosomes as Reconfigurable Therapeutic Systems. Trends Mol. Med. 2017, 23, 636–650. [Google Scholar] [CrossRef]
  541. Pullan, J.E.; Confeld, M.I.; Osborn, J.K.; Kim, J.; Sarkar, K.; Mallik, S. Exosomes as Drug Carriers for Cancer Therapy. Mol. Pharm. 2019, 16, 1789–1798. [Google Scholar] [CrossRef]
  542. Kim, H.; Jang, H.; Cho, H.; Choi, J.; Hwang, K.Y.; Choi, Y.; Kim, S.H.; Yang, Y. Recent Advances in Exosome-Based Drug Delivery for Cancer Therapy. Cancers 2021, 13, 4435. [Google Scholar] [CrossRef]
  543. Chen, L.; Wang, L.; Zhu, L.; Xu, Z.; Liu, Y.; Li, Z.; Zhou, J.; Luo, F. Exosomes as Drug Carriers in Anti-Cancer Therapy. Front. Cell Dev. Biol. 2022, 10, 728616. [Google Scholar] [CrossRef]
  544. Gorshkov, A.; Purvinsh, L.; Brodskaia, A.; Vasin, A. Exosomes as Natural Nanocarriers for RNA-Based Therapy and Prophylaxis. Nanomaterials 2022, 12, 524. [Google Scholar] [CrossRef] [PubMed]
  545. Li, J.; Stephens, E.; Zhang, Y. Chapter Nine—Extracellular Vesicles in Tumor Immunotherapy. In Systemic Drug Delivery Strategies; Amiji, M.M., Milane, L.S., Eds.; Academic Press: Cambridge, MA, USA, 2022; pp. 231–256. [Google Scholar] [CrossRef]
  546. Muhammad, S.A.; Mohammed, J.S.; Rabiu, S. Exosomes as Delivery Systems for Targeted Tumour Therapy: A Systematic Review and Meta-Analysis of In Vitro Studies. Pharm. Nanotechnol. 2023, 11, 93–104. [Google Scholar] [CrossRef]
  547. Wu, C.-Y.; Du, S.-L.; Zhang, J.; Liang, A.-L.; Liu, Y.-J. Exosomes and Breast Cancer: A Comprehensive Review of Novel Therapeutic Strategies from Diagnosis to Treatment. Cancer Gene Ther. 2017, 24, 6–12. [Google Scholar] [CrossRef] [PubMed]
  548. Allahverdiyev, A.M.; Parlar, E.; Dinparvar, S.; Bagirova, M.; Abamor, E.Ş. Current Aspects in Treatment of Breast Cancer Based of Nanodrug Delivery Systems and Future Prospects. Artif. Cells Nanomed. Biotechnol. 2018, 46, S755–S762. [Google Scholar] [CrossRef] [Green Version]
  549. Mughees, M.; Kumar, K.; Wajid, S. Exosome Vesicle as a Nano-Therapeutic Carrier for Breast Cancer. J. Drug Target. 2021, 29, 121–130. [Google Scholar] [CrossRef]
  550. Al-Humaidi, R.B.; Fayed, B.; Sharif, S.I.; Noreddin, A.; Soliman, S.S.M. Role of Exosomes in Breast Cancer Management: Evidence-Based Review. Curr. Cancer Drug Targets 2021, 21, 666–675. [Google Scholar] [CrossRef]
  551. Ha, D.; Yang, 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]
  552. Yang, B.; Chen, Y.; Shi, J. Exosome Biochemistry and Advanced Nanotechnology for Next-Generation Theranostic Platforms. Adv. Mater. 2019, 31, e1802896. [Google Scholar] [CrossRef] [PubMed]
  553. Rufino-Ramos, D.; Albuquerque, P.R.; Carmona, V.; Perfeito, R.; Nobre, R.J.; de Almeida, L.P. Extracellular Vesicles: Novel Promising Delivery Systems for Therapy of Brain Diseases. J. Control Release 2017, 262, 247–258. [Google Scholar] [CrossRef]
  554. Kutova, O.M.; Guryev, E.L.; Sokolova, E.A.; Alzeibak, R.; Balalaeva, I.V. Targeted Delivery to Tumors: Multidirectional Strategies to Improve Treatment Efficiency. Cancers 2019, 11, 68. [Google Scholar] [CrossRef] [Green Version]
  555. Smyth, T.J.; Redzic, J.S.; Graner, M.W.; Anchordoquy, T.J. Examination of the Specificity of Tumor Cell Derived Exosomes with Tumor Cells in Vitro. Biochim. Biophys. Acta BBA-Biomembr. 2014, 1838, 2954–2965. [Google Scholar] [CrossRef] [Green Version]
  556. Agrawal, A.K.; Aqil, F.; Jeyabalan, J.; Spencer, W.A.; Beck, J.; Gachuki, B.W.; Alhakeem, S.S.; Oben, K.; Munagala, R.; Bondada, S.; et al. Milk-Derived Exosomes for Oral Delivery of Paclitaxel. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 1627–1636. [Google Scholar] [CrossRef]
  557. Kandimalla, R.; Aqil, F.; Alhakeem, S.S.; Jeyabalan, J.; Tyagi, N.; Agrawal, A.; Yan, J.; Spencer, W.; Bondada, S.; Gupta, R.C. Targeted Oral Delivery of Paclitaxel Using Colostrum-Derived Exosomes. Cancers 2021, 13, 3700. [Google Scholar] [CrossRef]
  558. Kumar, D.N.; Chaudhuri, A.; Dehari, D.; Shekher, A.; Gupta, S.C.; Majumdar, S.; Krishnamurthy, S.; Singh, S.; Kumar, D.; Agrawal, A.K. Combination Therapy Comprising Paclitaxel and 5-Fluorouracil by Using Folic Acid Functionalized Bovine Milk Exosomes Improves the Therapeutic Efficacy against Breast Cancer. Life 2022, 12, 1143. [Google Scholar] [CrossRef]
  559. Zhang, P.; Zhang, L.; Qin, Z.; Hua, S.; Guo, Z.; Chu, C.; Lin, H.; Zhang, Y.; Li, W.; Zhang, X.; et al. Genetically Engineered Liposome-like Nanovesicles as Active Targeted Transport Platform. Adv. Mater. 2018, 30, 1705350. [Google Scholar] [CrossRef] [Green Version]
  560. Gomari, H.; Moghadam, M.F.; Soleimani, M.; Ghavami, M.; Khodashenas, S. Targeted Delivery of Doxorubicin to HER2 Positive Tumor Models. Int. J. Nanomed. 2019, 14, 5679–5690. [Google Scholar] [CrossRef] [Green Version]
  561. Pullan, J.; Dailey, K.; Bhallamudi, S.; Feng, L.; Alhalhooly, L.; Froberg, J.; Osborn, J.; Sarkar, K.; Molden, T.; Sathish, V.; et al. Modified Bovine Milk Exosomes for Doxorubicin Delivery to Triple-Negative Breast Cancer Cells. Acs. Appl. Biol. Mater. 2022, 5, 2163–2175. [Google Scholar] [CrossRef]
  562. Schindler, C.; Collinson, A.; Matthews, C.; Pointon, A.; Jenkinson, L.; Minter, R.R.; Vaughan, T.J.; Tigue, N.J. Exosomal Delivery of Doxorubicin Enables Rapid Cell Entry and Enhanced in Vitro Potency. PLoS ONE 2019, 14, e0214545. [Google Scholar] [CrossRef] [Green Version]
  563. Zhang, G.; Yang, X.; Su, X.; An, N.; Yang, F.; Li, X.; Jiang, Y.; Xing, Y. Understanding the Protective Role of Exosomes in Doxorubicin-Induced Cardiotoxicity. Oxidative Med. Cell Longev. 2022, 2022, 2852251. [Google Scholar] [CrossRef]
  564. Hadla, M.; Palazzolo, S.; Corona, G.; Caligiuri, I.; Canzonieri, V.; Toffoli, G.; Rizzolio, F. Exosomes Increase the Therapeutic Index of Doxorubicin in Breast and Ovarian Cancer Mouse Models. Nanomedicine 2016, 11, 2431–2441. [Google Scholar] [CrossRef] [PubMed]
  565. Ni, J.; Mi, Y.; Wang, B.; Zhu, Y.; Ding, Y.; Ding, Y.; Li, X. Naturally Equipped Urinary Exosomes Coated Poly (2−ethyl−2−oxazoline)−Poly (D, L−lactide) Nanocarriers for the Pre−Clinical Translation of Breast Cancer. Bioengineering 2022, 9, 363. [Google Scholar] [CrossRef] [PubMed]
  566. Kanchanapally, R.; Brown, K. Cancer Cell-Derived Exosomes as the Delivery Vehicle of Paclitaxel to Inhibit Cancer Cell Growth. J. Cancer Discov. 2022, 1, 49–58. [Google Scholar] [CrossRef]
  567. Liu, J.; Tang, Y.; Li, Y.; Hu, X.; Huang, S.; Xu, W.; Hao, X.; Zhou, M.; Wu, J.; Xiang, D. Paclitaxel-Loaded Hybrid Exosome for Targeted Chemotherapy of Triple-Negative Breast Cancer. Res. Sq. 2022. preprint. [Google Scholar] [CrossRef]
  568. Wang, K.; Ye, H.; Zhang, X.; Wang, X.; Yang, B.; Luo, C.; Zhao, Z.; Zhao, J.; Lu, Q.; Zhang, H.; et al. An Exosome-like Programmable-Bioactivating Paclitaxel Prodrug Nanoplatform for Enhanced Breast Cancer Metastasis Inhibition. Biomaterials 2020, 257, 120224. [Google Scholar] [CrossRef] [PubMed]
  569. Lin, Y.; Wu, J.; Gu, W.; Huang, Y.; Tong, Z.; Huang, L.; Tan, J. Exosome–Liposome Hybrid Nanoparticles Deliver CRISPR/Cas9 System in MSCs. Adv. Sci. 2018, 5, 1700611. [Google Scholar] [CrossRef] [Green Version]
  570. Naseri, Z.; Oskuee, R.K.; Jaafari, M.R.; Forouzandeh, M. Exosome-Mediated Delivery of Functionally Active MiRNA-142-3p Inhibitor Reduces Tumorigenicity of Breast Cancer in Vitro and in Vivo. Int. J. Nanomed. 2018, 13, 7727–7747. [Google Scholar] [CrossRef] [Green Version]
  571. Kim, H.; Rhee, W.J. Exosome-Mediated Let7c-5p Delivery for Breast Cancer Therapeutic Development. Biotechnol. Bioproc. E 2020, 25, 513–520. [Google Scholar] [CrossRef]
  572. 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] [Green Version]
  573. O’Brien, K.; Lowry, M.C.; Corcoran, C.; Martinez, V.G.; Daly, M.; Rani, S.; Gallagher, W.M.; Radomski, M.W.; MacLeod, R.A.F.; O’Driscoll, L. MiR-134 in Extracellular Vesicles Reduces Triple-Negative Breast Cancer Aggression and Increases Drug Sensitivity. Oncotarget 2015, 6, 32774–32789. [Google Scholar] [CrossRef] [Green Version]
  574. Shojaei, S.; Hashemi, S.M.; Ghanbarian, H.; Sharifi, K.; Salehi, M.; Mohammadi-Yeganeh, S. Delivery of MiR-381-3p Mimic by Mesenchymal Stem Cell-Derived Exosomes Inhibits Triple Negative Breast Cancer Aggressiveness; an In Vitro Study. Stem Cell Rev. Rep. 2021, 17, 1027–1038. [Google Scholar] [CrossRef]
  575. Sheykhhasan, M.; Kalhor, N.; Sheikholeslami, A.; Dolati, M.; Amini, E.; Fazaeli, H. Exosomes of Mesenchymal Stem Cells as a Proper Vehicle for Transfecting MiR-145 into the Breast Cancer Cell Line and Its Effect on Metastasis. Biomed. Res. Int. 2021, 2021, 5516078. [Google Scholar] [CrossRef]
  576. Gong, C.; Tian, J.; Wang, Z.; Gao, Y.; Wu, X.; Ding, X.; Qiang, L.; Li, G.; Han, Z.; Yuan, Y.; et al. Functional Exosome-Mediated Co-Delivery of Doxorubicin and Hydrophobically Modified MicroRNA 159 for Triple-Negative Breast Cancer Therapy. J. Nanobiotechnol. 2019, 17, 93. [Google Scholar] [CrossRef] [Green Version]
  577. Liu, X.; Zhang, G.; Yu, T.; He, J.; Liu, J.; Chai, X.; Zhao, G.; Yin, D.; Zhang, C. Exosomes Deliver LncRNA DARS-AS1 SiRNA to Inhibit Chronic Unpredictable Mild Stress-Induced TNBC Metastasis. Cancer Lett. 2022, 543, 215781. [Google Scholar] [CrossRef]
  578. Markov, O.; Oshchepkova, A.; Mironova, N. Immunotherapy Based on Dendritic Cell-Targeted/-Derived Extracellular Vesicles—A Novel Strategy for Enhancement of the Anti-Tumor Immune Response. Front. Pharmacol. 2019, 10, 1152. [Google Scholar] [CrossRef] [Green Version]
  579. 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]
  580. Zitvogel, L.; Regnault, A.; Lozier, A.; Wolfers, J.; Flament, C.; Tenza, D.; Ricciardi-Castagnoli, P.; Raposo, G.; Amigorena, S. Eradication of Established Murine Tumors Using a Novel Cell-Free Vaccine: Dendritic Cell Derived Exosomes. Nat. Med. 1998, 4, 594–600. [Google Scholar] [CrossRef]
  581. Pitt, J.M.; Charrier, M.; Viaud, S.; André, F.; Besse, B.; Chaput, N.; Zitvogel, L. Dendritic Cell–Derived Exosomes as Immunotherapies in the Fight against Cancer. J. Immunol. 2014, 193, 1006–1011. [Google Scholar] [CrossRef] [Green Version]
  582. Xie, Y.; Wu, J.; Xu, A.; Ahmeqd, S.; Sami, A.; Chibbar, R.; Freywald, A.; Zheng, C.; Xiang, J. Heterologous Human/Rat HER2-Specific Exosome-Targeted T Cell Vaccine Stimulates Potent Humoral and CTL Responses Leading to Enhanced Circumvention of HER2 Tolerance in Double Transgenic HLA-A2/HER2 Mice. Vaccine 2018, 36, 1414–1422. [Google Scholar] [CrossRef]
  583. Liu, H.; Chen, L.; Peng, Y.; Yu, S.; Liu, J.; Wu, L.; Zhang, L.; Wu, Q.; Chang, X.; Yu, X.; et al. Dendritic Cells Loaded with Tumor Derived Exosomes for Cancer Immunotherapy. Oncotarget 2017, 9, 2887–2894. [Google Scholar] [CrossRef] [Green Version]
  584. Huang, L.; Rong, Y.; Tang, X.; Yi, K.; Qi, P.; Hou, J.; Liu, W.; He, Y.; Gao, X.; Yuan, C.; et al. Engineered Exosomes as an in Situ DC-Primed Vaccine to Boost Antitumor Immunity in Breast Cancer. Mol. Cancer 2022, 21, 45. [Google Scholar] [CrossRef]
  585. Taghikhani, A.; Hassan, Z.M.; Ebrahimi, M.; Moazzeni, S. MicroRNA Modified Tumor-derived Exosomes as Novel Tools for Maturation of Dendritic Cells. J. Cell Physiol. 2019, 234, 9417–9427. [Google Scholar] [CrossRef]
  586. Jung, K.O.; Jo, H.; Yu, J.H.; Gambhir, S.S.; Pratx, G. Development and MPI Tracking of Novel Hypoxia-Targeted Theranostic Exosomes. Biomaterials 2018, 177, 139–148. [Google Scholar] [CrossRef] [PubMed]
  587. Limoni, S.K.; Moghadam, M.F.; Moazzeni, S.M.; Gomari, H.; Salimi, F. Engineered Exosomes for Targeted Transfer of SiRNA to HER2 Positive Breast Cancer Cells. Appl. Biochem. Biotech. 2019, 187, 352–364. [Google Scholar] [CrossRef]
  588. Cheng, Q.; Shi, X.; Han, M.; Smbatyan, G.; Lenz, H.-J.; Zhang, Y. Reprogramming Exosomes as Nanoscale Controllers of Cellular Immunity. J. Am. Chem. Soc. 2018, 140, 16413–16417. [Google Scholar] [CrossRef]
  589. Shi, X.; Cheng, Q.; Hou, T.; Han, M.; Smbatyan, G.; Lang, J.E.; Epstein, A.L.; Lenz, H.-J.; Zhang, Y. Genetically Engineered Cell-Derived Nanoparticles for Targeted Breast Cancer Immunotherapy. Mol. Ther. 2020, 28, 536–547. [Google Scholar] [CrossRef] [PubMed]
  590. Chen, C.; Zhao, S.; Karnad, A.; Freeman, J.W. The Biology and Role of CD44 in Cancer Progression: Therapeutic Implications. J. Hematol. Oncol. 2018, 11, 64. [Google Scholar] [CrossRef] [Green Version]
  591. Hong, Y.; Nam, G.; Koh, E.; Jeon, S.; Kim, G.B.; Jeong, C.; Kim, D.; Yang, Y.; Kim, I. Exosome as a Vehicle for Delivery of Membrane Protein Therapeutics, PH20, for Enhanced Tumor Penetration and Antitumor Efficacy. Adv. Funct. Mater. 2018, 28, 1703074. [Google Scholar] [CrossRef]
  592. Feng, C.; Xiong, Z.; Wang, C.; Xiao, W.; Xiao, H.; Xie, K.; Chen, K.; Liang, H.; Zhang, X.; Yang, H. Folic Acid-Modified Exosome-PH20 Enhances the Efficiency of Therapy via Modulation of the Tumor Microenvironment and Directly Inhibits Tumor Cell Metastasis. Bioact. Mater. 2021, 6, 963–974. [Google Scholar] [CrossRef] [PubMed]
  593. Li, D.; Yao, S.; Zhou, Z.; Shi, J.; Huang, Z.; Wu, Z. Hyaluronan Decoration of Milk Exosomes Directs Tumor-Specific Delivery of Doxorubicin. Carbohydr. Res. 2020, 493, 108032. [Google Scholar] [CrossRef] [PubMed]
  594. Piffoux, M.; Nicolás-Boluda, A.; Mulens-Arias, V.; Richard, S.; Rahmi, G.; Gazeau, F.; Wilhelm, C.; Silva, A.K.A. Extracellular Vesicles for Personalized Medicine: The Input of Physically Triggered Production, Loading and Theranostic Properties. Adv. Adv. Drug Deliv. Rev. 2018, 138, 247–258. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SEVs biogenesis and release. SEVs may have multiple origins. They can originate from plasma membrane budding, which leads to heterogeneous membranous medium-large vesicles (MLEVs) shedding. Small EVs (SEVs, exosomes) originate from the internal budding of plasma membranes giving rise to early endosomes. By complex maturating interactions with the Golgi apparatus (cargo-in/cargo-out), early endosomes become late ones. The membranes of late endosomes form intraluminal vesicles (ILVs), small cargos containing proteins from the plasma membrane, and Golgi as well as nucleic acids. Endosomal cargo sorting was performed through either ESCRT-dependent or -independent routes; the ESCRT complex being the key machinery of protein sorting into SEVs. ILVs are contained in multivesicular bodies (MVBs) that fuse with either plasma membrane (after Rab-driven docking), releasing SEVs in the extracellular space (through Snare complex assembly) or with lysosomes for further internal degradation.
Figure 1. SEVs biogenesis and release. SEVs may have multiple origins. They can originate from plasma membrane budding, which leads to heterogeneous membranous medium-large vesicles (MLEVs) shedding. Small EVs (SEVs, exosomes) originate from the internal budding of plasma membranes giving rise to early endosomes. By complex maturating interactions with the Golgi apparatus (cargo-in/cargo-out), early endosomes become late ones. The membranes of late endosomes form intraluminal vesicles (ILVs), small cargos containing proteins from the plasma membrane, and Golgi as well as nucleic acids. Endosomal cargo sorting was performed through either ESCRT-dependent or -independent routes; the ESCRT complex being the key machinery of protein sorting into SEVs. ILVs are contained in multivesicular bodies (MVBs) that fuse with either plasma membrane (after Rab-driven docking), releasing SEVs in the extracellular space (through Snare complex assembly) or with lysosomes for further internal degradation.
Ijms 24 07208 g001
Figure 2. Exosome membrane molecules and their cargo content. Small extracellular vesicles (SEVs) are nano-sized membrane vesicles released by a variety of cell types and are thought to play important roles in intercellular communications. SEVs contain many kinds of proteins, either cytosolic or plasma membrane ones. Transporters, receptors, and signaling proteins, but also enzymes, can be evidenced. Metabolites are also present as well as nucleic acids. Genomic and mitochondrial DNAs and multiple RNAs (mRNAs, miRNA, lncRNA, circRNA, etc.) can be detected. Through the horizontal transfer of these bioactive molecules, SEVS are emerging as local and systemic cell-to-cell mediators of oncogenic information.
Figure 2. Exosome membrane molecules and their cargo content. Small extracellular vesicles (SEVs) are nano-sized membrane vesicles released by a variety of cell types and are thought to play important roles in intercellular communications. SEVs contain many kinds of proteins, either cytosolic or plasma membrane ones. Transporters, receptors, and signaling proteins, but also enzymes, can be evidenced. Metabolites are also present as well as nucleic acids. Genomic and mitochondrial DNAs and multiple RNAs (mRNAs, miRNA, lncRNA, circRNA, etc.) can be detected. Through the horizontal transfer of these bioactive molecules, SEVS are emerging as local and systemic cell-to-cell mediators of oncogenic information.
Ijms 24 07208 g002
Figure 3. Bidirectional communication between tumor cells and their surrounding environment. The tumor microenvironment (TME) is a complex and dynamic network that includes normal breast (NBC), tumor (TC), cancer-associated fibroblasts (CAFs), mesenchymal stromal (MSC), immune (tumor-associated macrophages TAM), and endothelial cells (EC). TC can bidirectionally signal to each other through SEVs production. TC can produce SEVs that will regulate MSCs’, CAFs’, and TAMs’ differentiation and activity. MSCs as well as TCs can regulate ECs’ activity, especially in hypoxic situations. TAMs, CAFs, and ECs can cooperate to promote angiogenesis. An antitumor immune response is largely modulated by BC cells through either extracellular signaling molecules (cytokines, etc.) secretion or SEVs production and release. BC cells SEVs contain inhibiting or activating molecules that favor target cells expansion, mobilization, and recruitment (CD4+ T cells (LT4), Tregs, and MSCs), polarization and activation (tumor-associated macrophages TAMs M2), and block others (CD8+ T cells (LT8), dendritic cells (DC), and natural killer NK cells).
Figure 3. Bidirectional communication between tumor cells and their surrounding environment. The tumor microenvironment (TME) is a complex and dynamic network that includes normal breast (NBC), tumor (TC), cancer-associated fibroblasts (CAFs), mesenchymal stromal (MSC), immune (tumor-associated macrophages TAM), and endothelial cells (EC). TC can bidirectionally signal to each other through SEVs production. TC can produce SEVs that will regulate MSCs’, CAFs’, and TAMs’ differentiation and activity. MSCs as well as TCs can regulate ECs’ activity, especially in hypoxic situations. TAMs, CAFs, and ECs can cooperate to promote angiogenesis. An antitumor immune response is largely modulated by BC cells through either extracellular signaling molecules (cytokines, etc.) secretion or SEVs production and release. BC cells SEVs contain inhibiting or activating molecules that favor target cells expansion, mobilization, and recruitment (CD4+ T cells (LT4), Tregs, and MSCs), polarization and activation (tumor-associated macrophages TAMs M2), and block others (CD8+ T cells (LT8), dendritic cells (DC), and natural killer NK cells).
Ijms 24 07208 g003
Figure 4. SEVs cargo as relevant breast cancer biomarkers. Among all the molecules present in SEVs, only a subset (proteins, miRNAs, and LncRNAs) have been shown to be of potential clinical value in CRC detection, diagnosis, prognosis, and treatment response evaluation. All referenced markers were found to be differentially expressed in cancer patients and in healthy people: miRs are depicted in blue, LncRNAs in brown, CirRNAs in green, and proteins in pink. NACT: neoadjuvant chemotherapy.This figure was created with BioRender (DT2576K3SR agreement number).
Figure 4. SEVs cargo as relevant breast cancer biomarkers. Among all the molecules present in SEVs, only a subset (proteins, miRNAs, and LncRNAs) have been shown to be of potential clinical value in CRC detection, diagnosis, prognosis, and treatment response evaluation. All referenced markers were found to be differentially expressed in cancer patients and in healthy people: miRs are depicted in blue, LncRNAs in brown, CirRNAs in green, and proteins in pink. NACT: neoadjuvant chemotherapy.This figure was created with BioRender (DT2576K3SR agreement number).
Ijms 24 07208 g004
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

Loric, S.; Denis, J.A.; Desbene, C.; Sabbah, M.; Conti, M. Extracellular Vesicles in Breast Cancer: From Biology and Function to Clinical Diagnosis and Therapeutic Management. Int. J. Mol. Sci. 2023, 24, 7208. https://doi.org/10.3390/ijms24087208

AMA Style

Loric S, Denis JA, Desbene C, Sabbah M, Conti M. Extracellular Vesicles in Breast Cancer: From Biology and Function to Clinical Diagnosis and Therapeutic Management. International Journal of Molecular Sciences. 2023; 24(8):7208. https://doi.org/10.3390/ijms24087208

Chicago/Turabian Style

Loric, Sylvain, Jérôme Alexandre Denis, Cédric Desbene, Michèle Sabbah, and Marc Conti. 2023. "Extracellular Vesicles in Breast Cancer: From Biology and Function to Clinical Diagnosis and Therapeutic Management" International Journal of Molecular Sciences 24, no. 8: 7208. https://doi.org/10.3390/ijms24087208

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