Next Article in Journal
Synthesis, Structure, and Magnetic and Biological Properties of Copper(II) Complexes with 1,3,4-Thiadiazole Derivatives
Next Article in Special Issue
The Clinical and Molecular Profile of Lung Cancer Patients Harboring the TP53 R337H Germline Variant in a Brazilian Cancer Center: The Possible Mechanism of Carcinogenesis
Previous Article in Journal
PON-Fold: Prediction of Substitutions Affecting Protein Folding Rate
Previous Article in Special Issue
Aspirin Affects MDA-MB-231 Vesicle Production and Their Capacity to Induce Fibroblasts towards a Pro-Invasive State
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Deciphering the Functional Status of Breast Cancers through the Analysis of Their Extracellular Vesicles

by
Alexis Germán Murillo Carrasco
1,2,
Andreia Hanada Otake
1,2,
Janaina Macedo-da-Silva
3,
Veronica Feijoli Santiago
3,
Giuseppe Palmisano
3,4,
Luciana Nogueira de Sousa Andrade
1,2 and
Roger Chammas
1,2,*
1
Center for Translational Research in Oncology (LIM24), Instituto do Cancer do Estado de Sao Paulo (ICESP), Hospital das Clinicas da Faculdade de Medicina da Universidade de Sao Paulo (HCFMUSP), São Paulo 01246-000, Brazil
2
Comprehensive Center for Precision Oncology, Universidade de São Paulo, São Paulo 01246-000, Brazil
3
Departamento de Parasitologia, Instituto de Ciências Biomédicas, Universidade de São Paulo, São Paulo 05508-000, Brazil
4
School of Natural Sciences, Macquarie University, Macquarie Park, NSW 2109, Australia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(16), 13022; https://doi.org/10.3390/ijms241613022
Submission received: 30 June 2023 / Revised: 10 August 2023 / Accepted: 18 August 2023 / Published: 21 August 2023
(This article belongs to the Special Issue State-of-the-Art Molecular Oncology in Brazil 2.0)

Abstract

:
Breast cancer (BC) accounts for the highest incidence of tumor-related mortality among women worldwide, justifying the growing search for molecular tools for the early diagnosis and follow-up of BC patients under treatment. Circulating extracellular vesicles (EVs) are membranous nanocompartments produced by all human cells, including tumor cells. Since minimally invasive methods collect EVs, which represent reservoirs of signals for cell communication, these particles have attracted the interest of many researchers aiming to improve BC screening and treatment. Here, we analyzed the cargoes of BC-derived EVs, both proteins and nucleic acids, which yielded a comprehensive list of potential markers divided into four distinct categories, namely, (i) modulation of aggressiveness and growth; (ii) preparation of the pre-metastatic niche; (iii) epithelial-to-mesenchymal transition; and (iv) drug resistance phenotype, further classified according to their specificity and sensitivity as vesicular BC biomarkers. We discuss the therapeutic potential of and barriers to the clinical implementation of EV-based tests, including the heterogeneity of EVs and the available technologies for analyzing their content, to present a consistent, reproducible, and affordable set of markers for further evaluation.

1. Introduction

Breast cancer (BC) is currently the most commonly diagnosed type of cancer in the world. In 2020, this disease accounted for 25.5% of new cases among women, and it was responsible for 15.5% of the cases of death via cancer among women [1]. Due to these rates, it is important to continue searching for screening and therapeutic options. Cancer is a disease produced by specific tissue cells with uncontrolled growth. Tumor cells use cell-to-cell communication to spread through diverse strategies, invading lymph nodes or distant organs, evading growth suppressors, or inducing tumor-supporting angiogenesis [2].
Extracellular vesicles (EVs) represent one form of cell-to-cell communication through a range of different types of membranous nano- and microparticles (ranging from 30 nm to 5 µm in size), each with a specific biogenesis, cargo, and function [3]. EVs are involved in several biological processes such as cell signaling, cell proliferation, immune system modulation [4,5,6], and the production of amphisomes as a consequence of their fusion with autophagosomes [7]. EV proteins are markers for early diagnosis and are prognostic in several types of solid tumors. In lung cancer, EV leucine-rich alpha-2-glycoprotein 1 was reported to be up-regulated in urinary and lung tissue [8]. Moreover, ovarian cancer EVs are enriched in integrin, PI3 kinase, p53, Ras, and other proteins related to cancer development [9,10]. In gastrointestinal system cancers, the amount of exosome is increased in patients with colorectal cancer, and it is correlated with carcinoembryogenic antigen [11]. EVs also participate in cancer progression processes such as tumor invasion, growth, and metastasis [12,13]. For example, several studies have shown that EV protein content is associated with breast cancer cells, demonstrating its modulation during oncogenesis and tumor progression. EVs support cancer progression, signaling recipient cells’ motility and growth [14]. In breast cancer, increased EV biogenesis has been related to the activation of protease-activated receptor 2 (PAR2), which is induced by different factors, including coagulation factor-FVIIa and trypsin [15]. PAR2 is cleaved by trypsin, which triggers PIK3-dependent AKT phosphorylation. AKT’s phosphorylation induces Rab5a activation, resulting in actin polymerization, which contributes to microvesicle budding, cell migration, and invasion [16]. Moreover, three different signaling cascades contribute to EV production via PAR2: (1) actin polymerization via the sequential activation of P38, MK2, and HSP27; (2) ERK1/2 activation, which stimulates MLC2 via MLCK; and (3) MLC2 activation via ROCK-II independently of ERK1/2 activation. The activation of MLC2 and HSP27 is essential to actomyosin rearrangement and EV production [15].
The versatility attributed to EVs is due to their heterogeneity. EVs include at least three main groups of membranous particles: exosomes, ectosomes, and apoptotic bodies [3]. Exosomes (30–150 nm in size) are initially packaged by a cell membrane bud called a multivesicular body (MVB), whereas ectosomes (200 nm–5 µm in size) are produced by protrusions of the cell membrane shed by the cell. On the other hand, apoptotic bodies are formed via cell fragmentation during programmed cell death [3]. Therefore, these particles are classified according to their sizes and the expression of a signature of surface proteins, a consequence of their different biogenesis processes. Nevertheless, the classification of these EV subtypes neglects a fourth group of EVs with no expression of the minimal set of markers described for any of the three first groups. This group includes particles whose functions and surface proteins are still being discovered, for example, migrasomes [17]. Extracellular vesicle biology is a highly dynamic and relatively recent research field. The International Society of Extracellular Vesicles (ISEV) seeks to harmonize the studies in the field, and it continuously updates the markers, mechanisms, and technologies for classifying these EVs in documents published as Minimal Information for Studies of EVs (MISEV) [18]. As researchers are still adopting the newest edition of MISEV, which includes recommendations for characterizing EV subtypes, for this review, we did not focus on specific EV subgroups, except when the authors have defined their particles as belonging to one of these subgroups.
Furthermore, EVs have arisen as a growing field of research due to their main characteristics, namely, (i) a wide of sizes from nano- to microscale (30 nm–5 µm) enabling their classification into groups (small or large EVs) [18,19,20]; (ii) a great diversity of proteins present on the surface of their lipid bilayer membranes, generally related to the molecular signature of the cell of origin [20,21]; and (iii) rhe mixed composition of their lumen or cargo, consisting of messenger RNA (mRNA) [22], microRNA (miRNA) [23], long non-coding RNA (lncRNA), circular RNA (circRNA) [24], double-stranded DNA (dsDNA) [25], mitochondrial DNA (mtDNA) [26], and proteins [27].
These characteristics allow EVs to be present in human fluids such as blood, urine, or saliva as a relevant form of cell-free circulating material helping to provide special protection for their luminal cargo, increasing the latter’s stability and half-life [28]. Furthermore, there is currently a need to understand more about the biology behind these vesicles and their potential applications, stimulating the creation of new EV-related databases [29,30,31]. These databases aim to organize the information on EV profiling being continuously produced worldwide despite the technical challenges surrounding their isolation and characterization [32].
Since EVs are collected through minimally invasive methods and carry different types of information, these particles represent a suitable source of information that may lead to the improvement of breast cancer screening and treatment. However, we still face challenges related to (i) the requirement for the proper identification and quantification of EVs; (ii) the identification of breast-cancer-related markers in EVs; and (iii) strategies for taking advantage of these breast-cancer-related EVs to create new therapeutic options, among others (Figure 1).
Transcriptomic approaches are usually performed on EVs secreted by cell lines in vitro and also in body fluid samples, which could be used to select targets for regulation in tumors. Meanwhile, since exploratory proteomic approaches are conventionally performed using cell lines, we can use this approach to characterize EVs from a single cell type from a classic perspective and propose tools for recognizing tumor-derived EVs. Therefore, this review aims to support the development of strategies for overcoming these challenges by compiling proteomic and transcriptomic information previously described in relation to breast cancer-derived EVs.

2. Transcriptomic and Genomics of BC-Related EVs

In addition to proteins, nucleic acids are another important component of EVs. The nucleic acid profile of Evs is primarily characterized by RNA types such as mRNA [24,33,34,35], miRNA [36,37,38,39], circRNA [24,40], rRNA [41,42], siRNA [41], and lncRNA [24,43]. However, mtDNA [33,44] and viral DNA [45] were also described in EVs from breast cancer samples. Herein, we explored the transcriptomic and genomic profiles of BC-related EVs to search for common markers and specific signatures associated with breast tumor features.

2.1. miRNA Profile in BC-Related EVs

Extracellular RNA (exRNA) is found in body fluids and can be protected from RNases through binding to circulating proteins like Ago2 and being sorted into extracellular vesicles (EVs) [23,46]. Since it is known that cancer cells secrete heterogeneous populations of EVs and that these EVs indeed contain diverse species of functional RNAs [47], the concept of accessing tumor RNA using circulating EVs is gaining great attention in the oncological field.
Among all vesicular nucleic acids, miRNAs represent the largest (and most analyzed) group. So far, at least 128 vesicular miRNAs are being studied for breast cancer (Supplementary Table S1 [36,38,39,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123]). These studies comprise analyses conducted to characterize breast tumors as well as understand the tumorigenic process and propose candidate vesicular biomarkers for diagnosis and prognosis.
Despite the high diversity described for this type of RNA, EVs from breast cancer culture supernatants (Supplementary Table S1, Figure 2A) and liquid biopsies (Supplementary Table S1, Figure 2B) have been shown to have three miRNAs in common that can be potential markers of this tumor: hsa-miR-21, hsa-miR-122, and hsa-miR-1246.
Interestingly, the miRNAs hsa-miR-21 and hsa-miR-122 were found in EVs from cell lines belonging to all the main subtypes of breast cancer following the classification provided by Dai et al. (2017) [124]. Nevertheless, it is essential to note that this review includes studies presenting both exploratory (RNA-seq, microarray, barcoding hybridization, etc.) and targeted (qRT-PCR) experiments analyzing samples from different countries/ancestral groups. Targeted analyses have broadly described oncomiRs, e.g., hsa-miR-21 [125,126] (Figure 3). Subsequently, this miRNA was extensively tested in extracellular vesicles, with the results confirming their role as an inductor of tumor processes [36,37,38,65,72,110] despite the existence of a minority of studies with divergent results [90]. In addition, Arisan et al., 2021 demonstrated that vesicular hsa-miR-21 facilitates the expansion of breast cancer cells favoring the epithelial–mesenchymal transition (EMT) by targeting Wnt-11 [66]. Moreover, this miRNA can support bone metastasis via the inhibition of PDCD4 [36,110] and cancer-related thrombosis by blocking IL6R in endothelial progenitor cells [65]. Regarding their involvement in the tumor microenvironment, Donnarumma et al. (2017) suggested that hsa-miR-21 could be transferred from cancer-associated fibroblasts (CAFs) to breast cancer cells via exosomes [127]. The horizontal transfer from this miRNA creates a self-regulating cycle that drives the presence of hsa-miR-21 in tumor cells.
On the other hand, other microRNAs like hsa-let-7f, hsa-miR-221, and hsa-miR-770 have consistently been found in EVs from Triple-Negative BC (TNBC) cells (TNA and TNB groups, Figure 2A). Interestingly, Wei et al. (2014) showed that vesicular hsa-miR-221 induced tamoxifen resistance in MCF7 cells (Luminal A subtype) by repressing the expression of p27 and ERα [76]. However, if an estrogen-receptor-positive BC cell receives hsa-miR-221 from an estrogen-receptor-negative cell through EVs, it induces the repression of the receptor via the MAPK pathway [128] (Figure 3). In TNBC, this vesicular miRNA will induce an EMT by targeting PTEN while also stimulating the expression of mesenchymal markers such as Snail, Slug, N-cadherin, and vimentin [75]. These findings support the hypothesis that specific components of vesicular cargo can modulate the aggressiveness of cell lines.
Nevertheless, one should keep in mind that these findings raise some concerns about the similarity of cargoes between EVs produced by pure cell lines and those captured via liquid biopsies that are secreted by multiple and heterogeneous types of cells, including breast tumors and non-malignant cells such as platelets. Among all the vesicular miRNAs observed in body fluids (Figure 2B), there is still a notable difference between miRNAs found in EVs from plasma and serum. Although these sources can be separated from blood, they have been reported to include different EV-related compositions [129,130] and, according to this review, varied cargoes as well. Intriguingly, hsa-miR-21 and hsa-miR-188 were consistently found in EVs from the plasma and serum of BC patients (Figure 2B). In addition, the study conducted by Inubushi et al. (2020) described the presence of hsa-miR-21 in the tears of BC patients [118]. Since cancer research is moving toward analyzing liquid biopsy sources, these efforts are intended to improve the screening and selection of candidates who will undergo solid biopsies (which are highly invasive). More recently, the scientific community has been interested in classifying studies using systemic liquid biopsy sources (blood, plasma, or serum) and local liquid biopsies (tears or urine), with the latter group guaranteeing completely noninvasive tests. Therefore, further research in this field of study is warranted in order to assess its feasibility.
A frequently encountered difficulty in characterizing EVs from body fluids concerns the specific tumor subtype of each patient cohort (Figure 1). In our review, we identified hsa-miR-373, hsa-miR-296, hsa-miR-32, and hsa-miR-548d that were exclusively expressed in EVs from the serum of TNBC patients. In addition, hsa-miR-373 was found in the EVs of luminal BC patients [113]. A few years later, hsa-miR-27a/b, hsa-miR-335, hsa-miR-365, hsa-miR-376c, hsa-miR-382, hsa-miR-422a, hsa-miR-433, and hsa-miR-628 were found to be differentially expressed in plasma small vesicles (classified by the authors as exosomes) from either TNBC or HER2+ patients compared to healthy individuals. Moreover, the authors found that a specific set of microRNAs composed of hsa-miR-16, hsa-miR-328, and hsa-miR-660 was associated with lymph node status among HER2+ patients, indicating that vesicular microRNA can also be associated with clinicopathological characteristics [131]. Interestingly, based on the increased abundance of hsa-miR-150-5p, hsa-miR-576-3p, and hsa-miR-4665-5p in exosomes from the plasma of BC patients, Wu et al. (2020) were able to distinguish patients with recurrence from patients without disease recurrence [132]. Finally, in this review, we identified the following vesicular miRNAs present in patients from all BC subtypes: hsa-miR-122, hsa-miR-16, hsa-miR-93, hsa-miR-423, and hsa-miR-101. Tissue-specific alterations are associated with different expression profiles such as HER-2 overexpression [121], TNBC [91,133], or the stem-like profile among basal-like ductal carcinoma in situ (DCIS) tumors [134], demonstrating the possibility of modifying levels of vesicular miRNAs. Once these findings have been consistently reported by studies performed on different BC cohorts, we believe these miRNAs deserve further attention as potential BC biomarkers in EVs.
In addition, short non-coding RNAs, including miRNA and siRNA, are potential therapeutic options. In this case, it is possible to transfect nucleic acids in extracellular vesicles to induce transcriptional changes in recipient cells. For example, the suppressor miRNA hsa-let-7a was used to design anti-tumor extracellular vesicles for targeting EGFR [68] and c-Myc [79] to repress tumor growth. Subsequently, Bose et al. (2018) suggested editing tumor-derived EVs from SKBR3 BC cells so that they included Cy5-anti-miR-21 in their cargo. This technique allowed for the inhibition of the oncogenic hsa-miR-21 and the tracking of the biodistribution of these tumor-derived EVs [135]. Afterward, the same group developed an engineered version of Evs coated on polymeric nanocarriers and loaded with anti-miR-21 and anti-miR-10b [70]. The prototype was tested in an in vivo model of TNBC that demonstrated an improved biodistribution to the tumor and the ability to exert a synergic effect with doxorubicin to control tumor growth.
Following this concept, Taghikhani et al., 2018 modified the cargo of tumor-derived extracellular vesicles so that they included higher concentrations of hsa-miR-155, hsa-miR-142, and hsa-let-7i for the further treatment of immature dendritic cells [136]. Their results showed that vesicular hsa-let-7i, enhanced by hsa-miR-155 and hsa-miR-142, acted as a cell-free vaccine for cancer treatment by inducing the maturation of dendritic cells. Using this characteristic, some studies have proposed the modification of autologous BC-derived EVs including siS100A4 [137] and ASO-1537S [138]. siS100A4 targets S100A4, a previously reported gene for the EMT in cancer [139], whereas ASO-1537S is an antisense oligonucleotide against 1537S, a non-coding mitochondrial RNA identified in oncogenic processes [138]. After these insertions, modified EVs prevented metastasis in an in vivo model for BC, providing a non-immunogenic method of transferring information and controlling cancer.
On the other hand, research on suppressor miRNA that is lowly expressed in tumor-derived EVs also opens a new window for therapy development. According to Zhou et al. (2021), BC cells treated with EVs carrying hsa-miR-424-5p, a suppressor miRNA, inhibited their growth though the miRNA-driven repression of PD-L1 [140].
Amalgamating these precedents, researchers have proposed new tools for cell editing using EVs (Figure 1). The main idea is to select (or develop) specific subpopulations of EVs that we can use to target cancer cells or change their environments. Nevertheless, although the use of vesicular microRNAs for diagnosis and prognosis seems promising, no panels of EV microRNAs are available for BC diagnosis in a clinical setting due to a lack of standardization and reproducibility concerning EV isolation and microRNA identification, as discussed in the final section (Figure 1).

2.2. Other Regulatory Non-Coding RNAs in BC-Related EVs

As described above, vesicular miRNAs can inhibit mRNA expression in their target cell. Then, oncomiRs repress tumor suppressors, and suppressor miRNAs inhibit the expression of tumor promoters. Nevertheless, exploratory techniques such as RNA-seq have supported the discovery of competing endogenous RNA (ceRNA) in BC-derived EVs [24,43]. This type of RNA can attract miRNAs by competing with their mRNA targets, thereby adding a new level of EV-based expression regulation. Herein, a ceRNA can inhibit a suppressor miRNA in order to induce the overexpression of the mRNA target, which stimulates tumor growth.
Some lncRNAs, a type of ceRNA, have been reported to be overexpressed in EVs from BC tissues, cells, and human fluids. For example, the lncRNAs SNHG16 [141,142], HOTAIR [143], H19 [144], SNHG14 [145], AFAP1-AS1 [146], AGAP2-AS1 [147,148], BCRT1 [149], UCA1 [150], GS1-600G8.5 [151], and NEAT1 [152] were found to be overexpressed in BC-derived EVs (Figure 3). Moreover, they have demonstrated their ability to induce new features in recipient cells.
SNHG14, AFAP1-AS1, AGAP2-AS1, and BCRT1 have been proven to be carried by tumor-derived EVs in order to induce trastuzumab resistance in BC cells (Figure 3). Interestingly, they achieve the same outcome in different ways. SNHG14 affects the Bcl-2 apoptosis pathway [145]; AFAP1-AS1 stimulates the overproduction of HER2, leading to treatment ineffectiveness [146]; AGAP2-AS1 is packaged in vesicles by hnRNPA2B1 [148] to link ELAVL1 in the receptor cell in order to modulate autophagy activity via the transcription of ATG10 [147]; and BCRT1 inhibits hsa-miR-1303 in order to induce the expression of PTBP3, a tumor promoter and target of hsa-miR-1303 [149]. Moreover, other studies have shown that vesicular H19, NEAT1, and UCA1 lncRNAs induce resistance to doxorubicin [144], cisplatin/paclitaxel [152], and tamoxifen [150], respectively. In all cases, it is possible to revert treatment resistance by silencing the respective lncRNA.
Tumor growth and metastatic expansion are processes that involve many molecular regulations. Accordingly, the mRNA–miRNA–lncRNA axis is analyzed to understand cancer biology. This reseaerch explains how tumor-derived EVs help to disseminate a competitive advantage among cancer cells. However, in some cases, they can “educate” neighboring epithelial and immune cells in terms of facilitating tumor expansion. For example, vesicular SNHG16 lncRNA induces metastasis by blocking hsa-miR-892 in BC cells; this action naturally represses PPAPDC1A, a metastasis promoter [142]. But the same lncRNA is able to block hsa-miR-16-5p in γδT1 lymphocytes in order to induce SMAD5 expression, which upregulates CD73 at the membrane of these tumor-infiltrating lymphocytes [141]. Vesicular BCRT1 induces M2 polarization in macrophages, thereby boosting BC progression [149]. Also, the presence of vesicular GS1-600G8.5 was associated with the destruction of the blood–brain barrier, thus facilitating brain metastases in BC patients [151].
In the same way that vesicular lncRNA from BC cells can modify the metabolism of other cell types, EVs produced in other cell types affect tumor growth. As an example, BC-modulating vesicular HISLA and SNHG3 lncRNAs are expressed in tumor-associated macrophages (TAM) and cancer-associated fibroblasts (CAF), respectively. HISLA obstructs the interaction between PHD2 and HIF-1α that leads to the accumulation of HIF-1α, which regulates glycolytic metabolism in the tumor cell [153]. SNHG3 sponges hsa-miR-330-5p in tumor cells that positively regulate PKM, thereby increasing tumor growth [154]. However, the transcriptome from BC-derived EVs also includes suppressor lncRNAs such as XIST, whose repression promotes brain metastasis in BC patients [155]. The absence of XIST induces an accumulation of hsa-miR-503, which triggers M1-M2 polarization, thereby suppressing T-cell proliferation and, in turn, facilitating tumor expansion.
For the time being, we have limited information regarding circRNAs. Some descriptive studies have described circHIPK3 [156], circHIF1A [157], hsa-circRNA-000615 [158], hsa-circRNA-0005795, and hsa-circRNA-0088088 [159] as being overexpressed in BC-derived EVs playing vital roles in BC development (Figure 3). Interestingly, circHIPK3 was associated with trastuzumab resistance [156]. Moreover, circHIF1A was identified as an oncogenic RNA because it downregulates hsa-miR-149-5p levels [157]. Subsequently, low levels of hsa-miR-149-5p induce overexpression of NFIB, an oncogenic promoter that functions via p21 inhibition.

2.3. Messenger RNA (mRNA) in BC-Related EVs

Besides regulating RNA, some studies have shown that mRNA is present in EVs derived from or targeting BC cells. A few years ago, Conley and co-authors identified an mRNA signature of this disease in EVs from breast cancer patients at advanced stages [34]. Under the same context, Rodríguez et al. (2015) found elevated levels of eight metastasis-related mRNAs and 27 stemness-related mRNAs in exosomes from breast cancer patients with poor prognosis, suggesting that exo-mRNA can be used as a prognostic marker for BC patients [160].
Andreeva et al., 2021 have demonstrated that PIK3CA mutations can be transferred between BC cell lines via their EVs [33]. Using RNA and DNA sequencing, this study described the presence of mutated versions of the PIK3CA gene in EVs from BC cell lines. BC-derived EVs have been observed overexpressing some mRNAs. According to Rodríguez et al. (2015), vesicular levels of HTR7, NEUROD1, and HOXC6 mRNA are related to disease-free survival among BC patients (p < 0.05), while NANOG, HTR7, NEUROD1, and HOXC6 mRNA levels are associated with overall survival (p < 0.05) [160]. Moreover, vesicular ERCC1 mRNA could be indicative of tumor progression (Figure 3) via their association with metastatic BC (p = 0.03) [161]. Furthermore, mRNA transcripts packaged inside tumor-derived EVs can be used as a disease indicator through liquid biopsy as proposed by Hu et al. (2023) [162]. To achieve this, the authors propose isolating different subsets of EVs using mAbs targeting tumor epitopes. Subsequently, they demonstrated that the mRNA profiling results from PAM50 genes of specific subtypes of vesicles from breast cancer patients were 100% in accordance with the tumor tissue and could be used to predict BC subtypes using minimally invasive methods.
On the other hand, different cell types can transmit growth signals to BC cells via EVs. Yao et al., 2019 demonstrated that platelets from BC patients overexpress TPM3 mRNA and transfer it via EVs to BC cells in order to promote their growth [163].
However, it should be mentioned that these studies face challenges relating to the low abundance of vesicular RNAs and the lack of housekeeping RNA species with which to make comparisons between different samples. Indeed, these important issues push scientists in the field to discuss the standardization of and guidelines for EV RNA analysis (Figure 1). Within this context, Gorji-Bahri et al., 2021 have proposed using the YWHAZ gene alone or in combination with GAPDH and UBC as stable values for vesicular RNA quantification [164].

2.4. DNA Profiles in Breast Cancer-Derived EVs

Interestingly, besides the presence of PIK3CA mRNA in BC-derived EVs, Andreeva and co-authors also found fragments of mutated DNA in these nanocompartments [33]. Accordingly, one might assume that EVs can contribute to the discarding or transmission of pathogenic variants present in one cell to another. Ruhen et al. (2021) also demonstrated that tumors export Copy Number Variations (CNVs) through EVs. Although the transference of somatic mutations has been reported before in the form of circulating tumor DNA (ctDNA), some of these mutated copies are sheltered by EVs that allow for their direct detection using highly sensitive technologies [44].
Moreover, some studies have described the presence of mtDNA in BC-derived EVs. Fragments of mtDNA in BC-derived EVs are associated with invasiveness via the activation of Toll-like receptor 9 [165], chemoresistance in TNBC cells [166], and escape from metabolic quiescence in hormone-therapy-resistant BC via estrogen-receptor-independent oxidative phosphorylation [167]. As it has been shown, this is a hitherto unexplored field that needs more research, mainly to determine the genes responsible for mtDNA transmission that can serve as candidates for gene therapy (Figure 1). Although the PINK1 gene was identified as a driver of mtDNA-carrying EVs [165], there are likely other drivers of this feature.
In addition, non-human DNA was also reported in BC-derived EVs. This is the case of HPV-positive EVs isolated from the serum of BC patients [45]. There is an important finding that could extend liquid biopsy features for the determination of viral subtypes in EV from patients as demonstrated by De Carolis et al., 2019 [45]. Although vesicular RNA has shown better information delivery characteristics than DNA [168], EVs serve as important tools for cellular communication that evades immune control. Therefore, EVs, their molecular topography, and their cargo deserve further attention.
Overall, the nucleic acid composition of BC-derived vesicles is highly diverse. Furthermore, some of these components are related to tumor processes such as the modulation of aggressiveness, the EMT, the preparation of the metastatic niche, and treatment resistance. Although EVs cannot deliver many nucleic acids per particle [22] (Figure 1), we present an overview of the relevant nucleic-acid-based markers in Figure 3.

3. Proteome of EVs in Breast Cancer

3.1. Analysis of BC-Related EV Proteomes Using High-Throughput Technologies

In order to elucidate the role of EV-related proteins during breast cancer pathogenesis, we reviewed the literature to determine the proteomes of EVs associated with breast cancer. A total of 21 studies [14,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188] that analyzed EVs from human breast cancer cell lines and biofluids using high-throughput and exploratory technologies were selected. A total of 8312 proteins were reported. The proteins were divided into core level 1 (protein identification in 50–100% of studies), core level 2 (protein identification in 40–50% of the studies), and core level 3 (protein identification in less than 40% of the studies), and they are reported in Supplementary Table S2 [14,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188]. Six and twenty-six proteins were assigned to cores 1 and 2, respectively. Although these data indicate high variability in the qualitative proteome content of EVs associated with breast cancer (Figure 1), we selected a few studies that evaluated methodological strategies for isolating EVs from breast cancer cells and biofluids and characterized their protein content. As many proteomic results were produced using BC cell lines, we used the classification system developed by Dai et al. (2017) to standardize the nomenclature and disease subtypes potentially represented by these cell lines [124].
Harris et al., 2015 investigated the subcellular localization of proteins found in the EVs of three BC cell lines with different types of metastatic profiles, namely, MDA-MB-231 (Triple-negative B subtype), MCF-7 overexpressing Rab27b, and MCF-7 (Luminal A subtype), and found several cellular localizations (the majority were from cytoplasm/cytoskeleton and integral/peripheral membrane proteins) and a minor proportion of EV proteins from the Golgi apparatus, ER, or mitochondria [14].
Regarding the molecular functions of the EV proteins, the majority displayed a protein-binding function and hydrolase activity. Comparing the EV protein expression between the three cell lines, 85 proteins were differentially expressed, with the non-invasive breast cancer cell line (MCF-7, Luminal A subtype) containing several up-regulated proteins with a tumor suppression function. Among the up-regulated proteins in MCF-7 compared to MDA-MB-231 were Tetraspanin (CD63, CD81, CD9, and Tetraspanin-14), Adhesion proteins (Neural cell adhesion molecule 2, Integrin alpha-V, Integrin beta-5, Epithelial cell adhesion molecule, Alpha-Parvin, Claudin-3, Cadherin-1, CD99, CD276, and Clusterin), stress response proteins (Heat shock protein beta-1), Small GTPase superfamily proteins (Rho-related GTP-binding protein RhoB, GTPase NRas, Ras-related protein Rap-2c, and RacGTPase-activating protein 1), and Endosome Trafficking/Transport proteins (Multivesicular body subunit 12B, Vacuolar protein sorting-associated protein 28 homolog, Vacuolar protein sorting-associated protein 37B, and Multivesicular body subunit 12A) [14]. The up-regulated proteins in MDA-MB-231 and MCF-7 overexpressing Rab27b compared to wild-type MCF-7 were adhesion/motility/cytoskeleton proteins (Vimentin, Galectin-3-binding protein, Annexin A1, Plectin, and Filamin-B), cell-surface receptor proteins (Ephrin type-A receptor 2), and stress response proteins (Protein NDRG1, Stress-70 protein, mitochondrial, and heat shock protein HSP 90-beta). Many of these proteins that were dysregulated in both cell lines are associated with tumorigenesis and metastasis [14] (Figure 3). Another protein of the Rab family, Rab5a, was associated with high migration and invasion rates in MDA-MB-231 cells along with the promotion of high rates of vesicle shedding. Interestingly, this process could be externally regulated by activating protease-activated receptor 2 (PAR2), using trypsin or coagulation factor-FVIIa to promote the accumulation of Rab5a [15,16].
Kruger et al., 2014 performed a molecular characterization of EVs from MCF-7 and MDA-MB-231 BC cells [188]. Using sucrose gradient ultracentrifugation and LC-MS/MS, the authors identified 59 and 88 proteins in MCF-7 and MDA-MB-231 EVs, respectively. The identified proteins were grouped based on molecular functions, such as catalytic activity, protein transport, adhesion, and extracellular matrix activity. Both EVs isolated from MCF-7 and MDA-MB-231 supernatants presented differences in their protein content. The MCF-7-derived EVs showed a greater abundance of biomolecule-binding and protein transport activity, while the MDA-MB-231-derived EVs contained proteins with catalytic activity. Moreover, 24% of the EV proteins of MDA-MB-231 were extracellular matrix proteins. These findings can be correlated with the higher metastatic potential of MDA-MB-231 (Triple-negative B subtype) compared to MCF-7 cells (Luminal A subtype) [188].
This study reported several proteins identified in both MDA-MB-231 and MCF-7 EVs, such as proteins from the Annexin family (Figure 3), Histone H4, and Calmodulin. Furthermore, the common EV proteins participate in cellular growth, signaling pathways, epigenomic alterations, DNA and histone methyltransferase, and Akt pathway regulation [189]. Moreover, these processes can be related to the malignancy of cancer cells and the poor prognosis of breast cancer [190]. Heinemann et al., 2014 used the MDA-MB-231 breast cancer cell line to develop a simple and efficient method for purifying cancer exosomes [183]. A three-step filtration method was implemented as the first step of a normal flow prefiltration using a 0.1 µm filter to remove larger constituents such as intact cells and cell debris. In the second step, tangential flow filtration using a 500kDa membrane filter was performed in order to remove proteins, and the retentate was passed to the third step, namely, track-etch filtration, where low pressure is applied using a 0.1 µm filter to isolate exosomes and remove microvesicles. A total of 60 unique proteins were identified sequential filtration, including the exosome marker CD63. In the exosome content, membrane trafficking proteins and proteins associated with transcription regulation, signal transduction, and the epigenetic modulation of nucleic acids were identified [183]. Regarding the heterogeneity of the vesicle entities, it is important to mention that filter-based techniques usually require the use of an automatized mechanism in order to avoid user interference during vesicle collection (Figure 1). Nevertheless, these techniques seem to support a relevant volume of EV proteomic data, so their application to EVs from different cell lines is recommended.
Smyth et al., 2014 examined the exosome proteomes from MCF-7 and PC3, cell lines of mammary gland/breast, and prostate cancers [185]. The identified exosome-delivered MCF-7 proteins were Heat Shock Proteins 70, 90, and 27; CD9; Ras-related protein 13 (RAB-13 and others); Annexins 1, 2, 5, and 7; Pyruvate kinase (PKM); Alpha-enolase (ENO1); Glyceraldehyde-3-phosphate dehydrogenase (GAPDH); Glucose-6-phosphate 1-dehydrogenase (G6PD); and actin. The functions of the exosomal proteins were related to amino acid transport, protease inhibition, the cytoskeleton, GTPase, ATPase, transcription regulation, transduction, adhesion, and chaperoning. Regarding subcellular localization, most of the proteins were in the membrane and cytoplasm; however, proteins localized in the Golgi and nucleus were also identified in MCF-7 exosomes. Moreover, it was found that the exosomal lipid composition also contributes to facilitating adherence/internalization in recipient cells [185]. The current stage of analytical methods, together with the optimization of pre-analytical factors and computational strategies, allowed for a better understanding of the EVs lipid compositions [191]. A comparison of the lipid compositions between EVs and cells of origin [192] has shown that, in vitro, EVs are more enriched in cholesterol, sphingomyelin, glycosphingolipids, and phosphatidylserine compared to their origin cells, while phosphatidylcholine and phosphatidylinositol are more enriched in cells compared to exosomes. Quantitative analyses of oxysterols in exosomes released from breast cancer cells revealed that levels of 27-Hydroxycholesterol were higher in exosomes from MCF-7 cells compared to MDA-MB-231 and non-cancerous cells, showing a dependence of the levels of this exosomal lipid on the ER status of the cell of origin [193]. Another study compared the lipid profiles of EVs isolated from TNBC cell lines, namely, D3H2LN and D3H1, with high and low metastatic potential, respectively [194]. The exosomal levels of unsaturated diacylglycerols isolated from the highly metastatic cells were higher and stimulated angiogenesis through the protein kinase D signaling pathway.
The EVs isolated from the MCF10 and MDA-MB231 cell lines shared proteins such as regulators of cell death, membrane components, adhesion, and cell motility proteins. The EVs from MDA-MB-231 were enriched in proteins associated with transcriptional regulation, proteolysis, EV formation (annexin, LAMP-1), cell cycle (NUMA1), and adherence to extracellular matrices (EDIL3, collagen, vitronectin). These results showed differences in the EV protein content of invasive and non-invasive breast cancer cells, which can contribute to the metastatic process [171].
The EV proteomes of two tumorigenic breast cancer cell lines, invasive (SKBR3, Her2 subtype) and non-invasive (MCF-7), were compared to non-tumorigenic MCF-10a. In this study, the authors used a synthetic peptide (Vn96) with a high affinity for heat shock proteins (highly expressed in cancer cells) to isolate EVs. Heat shock proteins are present on the exosome surface, binding to Vn96 and facilitating EV recovery [195]. In total, 392 (SKBR3) and 301 (MCF-7) exosomal proteins were identified, and they were associated with membranes (19%) and the cell surface (12%). The functions of EV proteins from SKBR3 are related to metabolism (enolase, fatty acid synthase, phosphoglycerate kinase, fructose bisphosphatase 1, GAPDH, malate dehydrogenase, L-lactate dehydrogenase, aldehyde dehydrogenase, aldolase, triosephosphate isomerase, and glucosidase 2 subunit beta), binding (Selenium-binding protein 1, 60 kDa heat shock protein, Protein disulfide-isomerase, Lamin A/C, and Tumor protein D52), and assembly (Myosin-9, alpha-Actinin-4, Cytokeratin 16, Cytokeratin 18, Cytokeratin 8, and Cytokeratin 19). Although MCF-7 EV protein content displayed the same molecular function as SKBR3, different proteins were identified, namely, Aldolase, Pyruvate kinase, Tryptophan-tRNA ligase, Cathepsin D, Kynureninase, TER ATPase, Lactoferroxin-C, and Hexokinase-1, which are involved in metabolism, and HSP90-a, Agrin, and protein SET are related to binding. Cytokeratin 19, which was identified in both SKBR3 and MCF-7 EVs, has been suggested to be involved in more aggressive tumor proliferation, metastasis, and invasion [196] (Figure 3). Cytokeratin 8 and 18 were reported to be secreted cancer biomarkers when detected in the serum of breast cancer patients [197]. EV proteins from SKBR3 (Her2 subtype) and MCF-7 (Luminal A subtype) cells are involved in lipogenesis, glycolysis/gluconeogenesis, and tricarboxylic acid, corroborating the important role of altered metabolism in cancer. Moreover, glycolytic enzymes can protect cancer cells from stress by inhibiting apoptosis [198]. Taken together, these data show that EV cargo is associated with the progression, proliferation, metabolic control, and malignancy of breast cancer (Figure 3).
Koh et al., 2021 analyzed the modulation of the plasma EV proteome of breast cancer patients with and without cognitive impairment following anthracycline-based treatment [187]. Plasma from the early-stage breast cancer patients was collected longitudinally at three different time points (before the start of chemotherapy, 3 weeks after cycle 2 of chemotherapy, and 3 weeks of the last cycle of chemotherapy) from two groups (cognitive non-impaired and impaired). Circulating EVs were isolated through ultracentrifugation and analyzed using Tandem-Mass-Tag (TMT)-based quantitative proteomics. A total of 517 regulated proteins were identified in the cognitive impairment group, which was compared to a non-impairment group at three time points. Among these regulated proteins were EV markers, such as CD9, TSPAN14, and CD5L. Moreover, a down-regulation of galactosylceramidase from the plasma EV protein content was observed in the cognitive non-impaired group at timepoint T3 compared to T1. On the other hand, p2X purinoceptor and cofilin-1 were up-regulated at timepoint 3 compared to T1. In contrast, the EV content of the cognitive impairment group displayed a down-regulation of p2X purinoceptor, cofilin-1, nexilin, and ADAM10 at T3 compared to T1 [187]. In this study, the authors also analyzed N-glycosylation sites in the EV content from plasma samples. However, none of the regulated proteins were identified in the N-glycosylation analysis. However, CD5L, which is an EV marker, displayed two glycosylation sites in both groups (non-impaired and impaired). An altered N-glycosylation profile was identified in CD5L peptides, which presented decreased glycosylation in the non-impaired group. These findings contribute to our understanding of how different omics characteristics interact between EVs, their cells of origin, and their destination. Nevertheless, due to the heterogeneity of the tumor itself as well as the intra and interpatient heterogeneity (Figure 1), further studies would need to include more available information about the clinical statuses of these patients (age at diagnosis, molecular subtype, immunohistochemical profiles, etc.) in order to find patterns in the EV-related markers using large-scale cohort analyses.
The EVs released from cancer cells can regulate specific biological processes in recipient cells and modulate the central nervous system, thereby impairing the cognitive abilities of breast cancer patients [187]. Jordan et al., analyzed the differences in the protein cargo of EVs isolated from breast cancer cell lines and the plasma of breast cancer patients and healthy donors to identify the structural features that contribute to altered functional activities [171]. EVs from peripheral plasma samples of patients presented an increased degree of invasion of non-invasive breast cancer cells. Breast cancer plasma EVs displayed a unique proteome profile, with typical EV markers, such as CD9, CD81, CD63, and HSP70; Rab proteins; and clathrin. Moreover, the data were used in a comparison between the gene expression of MCF-10 treated with EVs obtained from breast cancer plasma patients, healthy donors, MDA-MB-231 cells, and untreated cells. The results displayed altered gene expression associated with angiogenesis, cell adhesion, and proliferation. In particular, it was confirmed that EVs released from aggressive breast cancer cells could stimulate the invasive behavior of non-invasive breast cancer cells. Moreover, the EV content from breast cancer patients can be used as a biomarker once these EVs contain a specific set of proteins [171].

3.2. Protein Markers in EVs Associated with Breast Cancer Features

Cancer-delivered EVs are a potential source of markers due to the fact that the cargo of EVs is enriched with molecules associated with tumor progression, metastasis, and invasion [199,200,201]. Rontogianni et al., analyzed the proteome content of EVs to discriminate cancer types and subtypes [176]. For this analysis, the authors used ten BC cell lines to identify a protein signature in EVs for use as a biomarker for breast cancer monitoring, diagnosis, and subtyping. Among the 4676 identified proteins, 14-3-3 proteins, integrins, annexin proteins, and cytoskeletal proteins, which are required for intermediate EV formation, were identified in all ten cancer cell lines. To discriminate the breast cancer subtypes, a total of 64 proteins were differentially expressed in TNBC compared to HER2+; these proteins are involved in angiogenesis (PLAU, ADAM9, and EPHA2), integrin-binding (ITGA5 and TIMP2), and cell motility (VIM and AXL). HER2+ EVs presented proteins associated with translation (EIFs), axon guidance (DNM2 and PIK3R1), and ERBB signaling (GRB7 and SHC1). Periostin, an extracellular matrix component, was described as a metastatic breast cancer biomarker identified in exosomes. It was observed that this protein was up-regulated in both murine and human cancer cell lines. Periostin has been reported to be highly expressed in several cancer types, including breast cancer [202], and it has been associated with osteoblast adhesion and cancer cell migration and is also involved in tumor angiogenesis [203]. Increased abundance of the Transient Receptor Potential Channel 5 (TrpC5) in EVs was observed in adriamycin-resistant human breast cancer cells (MCF-7/ADM) and was correlated with EV formation; additionally, adriamycin was found in EVs. The transference of TrpC5 to recipient cells mediated by EVs allows recipient cells to acquire TrpC5 and to P-glycoprotein multidrug transporter expression, conferring chemoresistance to non-resistance cells [204]. Considering the relevance of studies involving several BC cell lines, there are two main factors that must be considered. First, previous studies on tumor-derived EVs have determined that culture conditions can affect the omics components of EV cargo [205]. Second, despite the above, it is necessary to have a reliable collection of vesicular omics data from BC cell lines to find consistent markers for further research (Figure 1).
Regarding the use of liquid biopsies in relation to metastatic breast cancer, in one study, EV proteins were monitored to identify potential biomarkers for prognostics using a thermophoretic aptasensor. A total of 286 plasma samples from metastatic patients, non-metastatic patients, and healthy donors were submitted to a thermophoretic aptasensor to obtain a profile of eight EV markers (Figure 3): CA 15-3, CA 125, carcinoembryonic antigen (CEA), human epidermal growth factor receptor 2 (HER2), epidermal growth factor receptor (EGFR), prostate-specific membrane antigen (PSMA), epithelial cell adhesion molecule (EpCAM), and vascular endothelial growth factor (VEGF). These markers are involved in survival and migration, cell proliferation, metastasis, invasion, cell stemness, vascular permeability, and angiogenesis [200]. In an independent study, sera-derived exosomes from breast cancer patients were used to characterize CD24 and EpCAM as markers for exosomes (Figure 3). Western-blotting analysis was performed to identify exosomal markers in breast cancer samples. EpCAM and CD24 were identified in exosomes isolated from ascites ovarian cancer patients. However, only CD24 was detected in the EVs isolated from the serum of breast cancer patients. The absence of EpCAM in the serum-derived exosomes from breast cancer can be explained by the cleavage of EpCAM by serum metalloproteinases and the fact that both CD24 and EpCAM are not present in the same EV populations, which was confirmed in this study after the immunoaffinity purification of exosomes using anti-EpCAM beads [206]. Taken together, these studies show the importance of EV proteins as breast cancer markers for clinical use in order to diagnose, stratify, and monitor breast cancer progression.

3.3. EV Protein Glycosylation in Breast Cancer

Glycoconjugates such as glycoproteins functionally decorate the cell surface and are released in the extracellular milieu, playing an essential role in cell communication with the extracellular environment in several pathophysiological conditions, including cancer [207,208]. Any alteration in glycosylation levels due to the variation in the expression of glycosyltransferases and glycosidases [209], the dysregulation of chaperone activity [210], different levels of substrates, and alterations in nucleotide sugar transporter levels are diagnostic signs in cancer [211]. Aberrant glycosylation in tumor cells influences angiogenesis, cell proliferation, invasion, and metastatic processes [212,213,214,215]. The surface of an EV is enriched with glycoproteins, such as the Tetraspanin CD63, which is known to regulate cancer malignancy, and glycolipids [216]. EV glycosylation has been reported in several cancer types, including breast cancer. A list of 15 studies on EV glycosylation in breast cancer was reported by Macedo-Silva et al. (2021) [217], which is provided herein in Supplementary Table S3 [169,170,172,182,183,184,185,190,218,219,220,221,222,223,224].
Nishida-Aoki et al. (2020) developed an EV glycosylation profile from breast cancer cells with brain metastasis tropism (BMD2a) comparing lymph node-metastatic (MDA-MB-231-luc-D3H2LN) and primary tumor-derived TNBC cell lines (MDA-MB-231-luc-D3H1) using lectin blotting [225]. The data showed a reduced level of sialic acid and a specific glycosylation pattern associated with galactose, GalNAc, lactose, and GalNAcα1−3GalNAc in BMD2a EVs. This study concluded that the glycosylation of EVs exerted an inhibitory effect on EV uptake [225]. This evidence shows that EV uptake depends on glycosylation profiles, as the authors showed that the cessation of O-glycosylation increases tumor EV accumulation in the lungs; meanwhile, N-glycosylation inhibition does not alter tumor EV biodistribution. It appears that the specific glycosylation profile in tumor EVs highlights their importance in breast cancer aggressiveness. For instance, EVs secreted by TNBC cells present higher levels of sialylation when compared to parental cells [226], and this process could interfere with EVs uptake. Sialic acid is negatively charged; its removal from the surfaces of EVs can promote the interaction of EVs and recipient cells through the uncovering of carbohydrate ligands that bind to lectins and promote the uptake of EV by recipient cells. Additionally, EV cargo contains glycosylated proteins that can interfere with EV–cell interaction. Cytokines binding to surface glycosaminoglycan side chains of proteoglycans decorate EVs isolated from breast cancer patients, altering EV biodistribution and leading to a higher metastatic burden [227]. The structure of bisecting GlcNAc glycan (β1,4-linked GlcNAc at the β-mannose residue core) on target proteins (EGFR and integrins) plays a role in cell metastasis and adhesion. The role of bisecting GlcNAc of EVs in breast cancer cells was described by Tan et al. (2020) [228]. Some proteins were described as being markers of EVs from tumor cells in breast cancer, such as extracellular matrix metalloproteinase inducer (EMMPRIN). A higher abundance of glycosylated EMMPRIN was identified in EVs from metastatic breast cancer patients. PNGase F treatment resulted in EVs’ deglycosylation and inhibited EV-induced invasion [222]. Somehow, glycans being components of EV cargo seems to dictate cell function. ST6Gal1-mediated sialylation modulates cell surface receptor function, promoting cell proliferation and invasion. Hait et al., showed that ST6Gal1 RNA transcripts and protein expression were variable and heterogeneous in a panel of TNBC and ER+ cell lines. However, EVs containing ST6Gal1 potentiate aggressive cancer cell growth, proliferation, and invasion in cells containing low amounts of endogenous ST6Gal1 [229]. In a heterogeneous tumor microenvironment, EVs act as important mediators of tumor progression, dictating EV organ tropism and metastasis. CA15-3/CA27.29 epitopes from Mucin 1 were reported to be tumor markers for breast cancer diagnosis. Additionally, alterations in Mucin 1 glycosylation profiles were observed in several cancer types, including human breast cancer, and in exosomes derived from luminal A breast carcinoma cells MCF-7 [170,230]. Extracellular matrix protein nephronectin (NPNT) could be present in a truncated or a glycosylated form in EVs secreted by mouse breast carcinoma cells [231]. N-glycosylation sites in integrin β1 modulate cell migration and the adhesion of MCF7 cells (luminal A), activating focal adhesion kinase (FAK) signaling, and these sites are crucial to integrin activity in EVs [232]. Additionally, exosomes released by MCF-7 cells after irradiation (luminal A type) increase concentrations of the enzyme GalNAc-T6, which is important for promoting epithelial–mesenchymal transition [233]. Furthermore, there is still a gap concerning whether alterations in EV glycosylation are due to changes in glycoprotein levels or changes in the expression/activity levels of glycan-modifying enzymes (Figure 1).

3.4. EV Proteome during the Metastatic Process in Breast Cancer

A screening of the literature regarding EV proteins identified in human and other species’ breast cancer cells, tissues, and biofluids, focusing on the TNBC and metastatic cases, allowed us to retrieve 22 studies and a total of 7265 proteins [14,170,171,173,175,176,177,178,179,181,184,186,187,188,221,234,235,236,237,238,239,240] (Supplementary Table S4). Interestingly, we observed that one set of proteins was identified more frequently (Figure 4A), with HLA-A, HLA-B, A2M, ACTB, ALDOA, ANXA2, and FN1 identified in at least 13 of 22 studies that explored the vesicular proteome in breast cancer.
Among the most frequent proteins, those located in membranes (plasma and organellar) and the cytoskeleton were particularly prominent (Figure 4A). On the other hand, after evaluating the subcellular localization of all the proteins identified at least twice in the EVs derived from breast cancer cells, it was determined that the proteasome complex, cell–cell adhesion junctions, focal adhesions, extracellular matrix, extracellular exosome, and membranes were also relevant (Figure 4B). It is possible to identify vesicular proteins that are involved in different phases of the cell cycle and the p53-linked response (Figure 4C), reinforcing the role of EVs in metastatic processes and breast cancer progression.
In fact, studies have demonstrated the crucial role of EVs in pro-metastatic processes, wherein a cancer cell’s communication with its local environment is necessary to initiate growth and invasion through the transfer of molecules, which include mRNAs, microRNAs, and proteins [235]. Different studies have revealed the vesicular proteins associated with metastatic events (Table 1), and among those listed, ADAM10 (four studies), VIM (three studies), and ANXA6 (two studies) stand out. An important finding was that exosomes expressing these proteins play important roles in TNBC (Figure 3).
It has been identified that the levels of the CD151 protein are increased in serum exosomes from TNBC patients, and this protein has been shown to promote proliferation, migration, and invasion by modulating the activities of laminin-binding integrins α3β1, α6β4, and α6β1 [175]. An increase in CD151 abundance was also verified in a TNBC cell model (MDA-MB-231) [181]. These results indicate that exosomal CD151 can be used as a biomarker to distinguish TNBC patients from healthy individuals.
The data gathered in Table 1 reveal the importance of other vesicular proteins in TNBC, such as VIM (three studies) and ANXA6, EDIL3, EPHA2, FLNA, FLNB, and PLAU (two studies). It has been observed that the VIM protein and its gene play important roles in controlling migration and invasion in different types of cancer [242,243]; here, it is highlighted that its location in EVs can be associated with the aggressiveness of cancer. Based on the collected data, it is possible to select possible markers of EVs in breast cancer. In addition to conventional markers such as CD9, CD81, and CD83, the proteins HLA-A, HLA-B, HSPA5, A2M, ACTB, ALDOA, ANXA2, FN1, ANXA5, GSN, MYH9, PGK1, and ACTN1 are also often identified in EVs in cells and breast cancer patients at different stages. Therefore, these markers could be used to purify EVs potentially related to highly aggressive tumor subpopulations and study them separately (Figure 1).

4. Current Status of BC-Related Biomarkers in EVs

EVs are natural sources of cancer biomarkers. In Table 2, we summarize potential markers described as vesicular cargo.
Considering that miRNAs are the most-studied components in the vesicular context, most putative markers for BC-related EVs depend on these non-coding RNA sequences. Herein, we found potential biomarkers for diagnosis, classification, prognosis, and the prediction of treatment response in relation to BC. Despite differences in the methods used for EV isolation and target selection, all area-under-the-curve (AUC) values shown for the diagnosis of BC were higher than 0.7.
Notably, vesicular hsa-miR-21 was identified as a relevant candidate for diagnosing BC patients in different cohorts [38,39,106]. According to some studies, its performance could be improved by estimating multi-miRNA scores [106] or by combining the levels of this miRNA with MMP1/CD63 protein quantification [37]. In addition, vesicular hsa-miR-421 levels demonstrated the potential to discriminate BC patients from healthy individuals, even those with early stages of the disease [115]. Furthermore, the analysis of vesicular miRNA levels revealed a dual function of some miRNAs. For example, hsa-miR-1246 [38,98] and hsa-miR-142-5p [244] allow for the diagnosis of BC patients and contribute to estimating the treatment response or classifying BC patients, respectively.
Moreover, other kinds of RNA markers (lncRNA, mRNAs, and circRNA) are currently being studied to characterize their features in EVs from BC patients (Table 2) and accumulate evidence to improve current classification systems. It is important to note that classification features can be enhanced if we specially analyze tumor-derived EVs. Although we still lack an exclusive tumor-cell-derived BC marker for EVs, some tumor markers, such as CD49f or the epithelial cell-specific marker (EpCAM), have shown promising results in this regard [106].

5. EVs in Oncological Precision Medicine and Treatment: Promising Avenues and Open Questions

There is no doubt that EVs are an important component of the tumor microenvironment, and a great deal of advancement has been made in understanding how these subcellular nanocompartments affect tumor progression and dissemination. Since tumor-derived EVs can be found in all biological fluids and EV cargo contains genetic information from parental cells, they have been considered potential candidates for liquid biopsies. One approach for specifically investigating circulating EVs derived from tumor cells is based on tumor surface markers that are present in EV membranes. In this context, glycoconjugates have recently attracted attention. Specific antibodies or lectins that recognize the aberrant glycosylation patterns of tumors serve as potential tools that can be used to isolate circulating-tumor-derived-EVs, allowing for the investigation of vesicular components such as RNA and proteins with clinical significance as biomarkers.
Nevertheless, despite EVs’ promising applicability and the advances in their use as a liquid biopsy tool, there are still some pitfalls that should be addressed by researchers. Technical challenges concerning effective and selective EV isolation, as well as EV nomenclature, still require better standardization to allow for the comparison of data between different groups. Additionally, we should certainly consider additional factors or biases that affect the collection of EVs from biological fluids like handling, storage, the use of anticoagulants, the volume of blood drawn, age, sex, ancestry groups, disease state (in addition to other comorbidities), medication, circadian rhythm, body mass index, hydration, and dietary status. Indeed, all these issues should be considered when data are interpreted. In addition, another layer of complexity can be added with respect to the methods for the quantitative analysis of EV cargo. First, the quantification of small amounts of RNA and, to a certain extent, protein are non-trivial. In particular, the required input of RNA often does not match the minimal amount recommended for different downstream procedures, leading to a poor signal-to-background ratio. Furthermore, the lack of normalization methods for the quantitative analysis of nucleic acids, proteins, and other biological molecules constitutes yet another gap in the field. Therefore, the scientific community is engaging in extensive discussions and research to establish pipelines to fill these gaps.
Despite the many issues mentioned above, better biomarkers for early detection and predicting therapeutic response are still needed in order to provide a more favorable clinical outcome. As for any new, growing field in science, some technical obstacles, such as those mentioned above, are still a reality; however, due to the number of circulating EVs and their nature as intrinsic sources of biological molecules for modulating cell phenotypes, we believe that circulating EVs hold a great deal of potential in the clinical setting since they represent a method for tracking cancer-associated changes in a personalized manner. Moreover, as the field advances toward overcoming these limitations, the information gleaned from EVs will certainly represent an interesting approach not only to understanding the clinical evolution of BC but also to revealing new routes for clinical intervention.
From the perspective of therapy, EVs have been considered as an important tool with which to improve cancer treatment. Since EVs are natural carriers, some strategies focus on loading EVs with therapeutic components such as chemotherapeutic drugs, siRNAs, peptides, and proteins, presenting significantly enhanced anti-tumor effects [251]. Different methods can be used for EV loading, including electroporation, extrusion, sonication, freeze–thawing, the use of pH gradients, ultrasound, the use of streptomycin O, and hybridization. Another alternative that has been used is the anchorage of a drug on an EV’s surface, which can be achieved via click chemistry, to modify the surface of EVs (reviewed in [252]).
Another approach of EV engineering is more indirect and involves the manipulation of the cellular EV source. In this approach, producing cells are incubated with therapeutic agents to increase the loading of these drugs into EVs. Furthermore, the producing cells can be transfected or infected with plasmids or viral vectors to overexpress specific molecules in secreted EVs. The presence of these biomolecules on an EV’s surface can be a promising strategy for tumor targeting and specific delivery, thereby improving therapeutic efficacy and reducing toxicity [253]. Nevertheless, it should be mentioned that there is not a gold-standard method for modifying EVs, and each one has its own advantages and disadvantages. Nonetheless, there are a few ongoing clinical trials using engineered EVs for cancer treatment (Clinical Trials, NCT03608631 and NCT01294072).

6. Conclusions

EVs play a substantial role in BC initiation and progression, impacting the natural history of the disease. Tumor-derived EVs are biological nanostructures that carry important genetic information from tumor cells and can be found in all biological fluids, representing a natural source of diagnostic and prognostic biomarkers in cancer. Within this context, finding a set of biomarkers in circulating EVs able to (1) discriminate between healthy individuals and BC patients, (2) discriminate between different molecular subtypes of BC, and (3) determine the probability of disease recurrence would represent a game changer in the field that could pave the way for precision oncology medicine. In addition, in the therapeutic setting, EV engineering can greatly improve the specificity and effectiveness of tumor therapy, as demonstrated in pre-clinical studies. However, further efforts are still necessary to better standardize technical procedures for EVs’ isolation and cargo identification along with the establishment of reliable assays with which to evaluate the therapeutic potential of EVs in clinical practice.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241613022/s1.

Author Contributions

Conceptualization, A.G.M.C., A.H.O., L.N.d.S.A. and R.C.; investigation, A.G.M.C., A.H.O., J.M.-d.-S., V.F.S., G.P., L.N.d.S.A. and R.C.; writing—original draft preparation, A.G.M.C., A.H.O., J.M.-d.-S., V.F.S., G.P., L.N.d.S.A. and R.C.; writing—review and editing, A.G.M.C., A.H.O., J.M.-d.-S., V.F.S., G.P., L.N.d.S.A. and R.C.; supervision, A.H.O., G.P., L.N.d.S.A. and R.C.; funding acquisition, R.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grants #2019/05583-0 (A.G.M.C. and R.C.), #2021/00140-3 (J.M.-d.-S.), #2018/18257-1 (G.P.), #2018/15549-1 (G.P.), and #2020/04923-0 (G.P.) from the São Paulo Research Foundation (FAPESP); CNPq grant #870219/1997-9 (V.F.) and grant #305239/2022-8 (R.C.); and the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior–Brasil (CAPES)–Finance Code 001 (A.G.M.C. and R.C.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

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] [PubMed]
  2. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed]
  3. Dixson, A.C.; Dawson, T.R.; Di Vizio, D.; Weaver, A.M. Context-specific regulation of extracellular vesicle biogenesis and cargo selection. Nat. Rev. Mol. Cell Biol. 2023, 24, 454–476. [Google Scholar] [CrossRef] [PubMed]
  4. Bobrie, A.; Colombo, M.; Raposo, G.; Théry, C. Exosome Secretion: Molecular Mechanisms and Roles in Immune Responses. Traffic 2011, 12, 1659–1668. [Google Scholar] [CrossRef] [PubMed]
  5. Mathivanan, S.; Ji, H.; Simpson, R.J. Exosomes: Extracellular organelles important in intercellular communication. J. Proteom. 2010, 73, 1907–1920. [Google Scholar] [CrossRef]
  6. Théry, C. Exosomes: Secreted vesicles and intercellular communications. F1000 Biol. Rep. 2011, 3, 15. [Google Scholar] [CrossRef]
  7. Gross, J.C.; Parbin, S. Crossroads of the endosomal machinery: Multivesicular bodies, small extracellular vesicles and autophagy. Trillium. Exctracellular. Vesicles 2020, 2, 48–53. [Google Scholar] [CrossRef]
  8. Li, Y.; Zhang, Y.; Qiu, F.; Qiu, Z. Proteomic identification of exosomal LRG1: A potential urinary biomarker for detecting NSCLC. Electrophoresis 2011, 32, 1976–1983. [Google Scholar] [CrossRef]
  9. Liang, B.; Peng, P.; Chen, S.; Li, L.; Zhang, M.; Cao, D.; Yang, J.; Li, H.; Gui, T.; Li, X.; et al. Characterization and proteomic analysis of ovarian cancer-derived exosomes. J. Proteom. 2013, 80, 171–182. [Google Scholar] [CrossRef]
  10. Sinha, A.; Ignatchenko, V.; Ignatchenko, A.; Mejia-Guerrero, S.; Kislinger, T. In-depth proteomic analyses of ovarian cancer cell line exosomes reveals differential enrichment of functional categories compared to the NCI 60 proteome. Biochem. Biophys. Res. Commun. 2014, 445, 694–701. [Google Scholar] [CrossRef]
  11. Silva, J.; Garcia, V.; Rodriguez, M.; Compte, M.; Cisneros, E.; Veguillas, P.; Garcia, J.M.; Dominguez, G.; Campos-Martin, Y.; Cuevas, J.; et al. Analysis of exosome release and its prognostic value in human colorectal cancer. Genes Chromosom. Cancer 2012, 51, 409–418. [Google Scholar] [CrossRef] [PubMed]
  12. Park, J.E.; Sen Tan, H.; Datta, A.; Lai, R.C.; Zhang, H.; Meng, W.; Lim, S.K.; Sze, S.K. Hypoxic tumor cell modulates its microenvironment to enhance angiogenic and metastatic potential by secretion of proteins and exosomes. Mol. Cell. Proteom. 2010, 9, 1085–1099. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, X.; Yuan, X.; Shi, H.; Wu, L.; Qian, H.; Xu, W. Exosomes in cancer: Small particle, big player. J. Hematol. Oncol. 2015, 8, 83. [Google Scholar] [CrossRef] [PubMed]
  14. Harris, D.A.; Patel, S.H.; Gucek, M.; Hendrix, A.; Westbroek, W.; Taraska, J.W. Exosomes Released from Breast Cancer Carcinomas Stimulate Cell Movement. PLoS ONE 2015, 10, e0117495. [Google Scholar] [CrossRef] [PubMed]
  15. Das, K.; Prasad, R.; Singh, A.; Bhattacharya, A.; Roy, A.; Mallik, S.; Mukherjee, A.; Sen, P. Protease-activated receptor 2 promotes actomyosin dependent transforming microvesicles generation from human breast cancer. Mol. Carcinog. 2018, 57, 1707–1722. [Google Scholar] [CrossRef]
  16. Das, K.; Prasad, R.; Roy, S.; Mukherjee, A.; Sen, P. The Protease Activated Receptor2 Promotes Rab5a Mediated Generation of Pro-metastatic Microvesicles. Sci. Rep. 2018, 8, 7357. [Google Scholar] [CrossRef]
  17. Ma, L.; Li, Y.; Peng, J.; Wu, D.; Zhao, X.; Cui, Y.; Chen, L.; Yan, X.; Du, Y.; Yu, L. Discovery of the migrasome, an organelle mediating release of cytoplasmic contents during cell migration. Cell Res. 2015, 25, 24–38. [Google Scholar] [CrossRef]
  18. 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]
  19. Alberro, A.; Iparraguirre, L.; Fernandes, A.; Otaegui, D. Extracellular Vesicles in Blood: Sources, Effects, and Applications. Int. J. Mol. Sci. 2021, 22, 8163. [Google Scholar] [CrossRef]
  20. Martins, Á.M.; Ramos, C.C.; Freitas, D.; Reis, C.A. Glycosylation of Cancer Extracellular Vesicles: Capture Strategies, Functional Roles and Potential Clinical Applications. Cells 2021, 10, 109. [Google Scholar] [CrossRef]
  21. Li, Y.; He, X.; Li, Q.; Lai, H.; Zhang, H.; Hu, Z.; Li, Y.; Huang, S. EV-origin: Enumerating the tissue-cellular origin of circulating extracellular vesicles using exLR profile. Comput. Struct. Biotechnol. J. 2020, 18, 2851–2859. [Google Scholar] [CrossRef] [PubMed]
  22. 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] [PubMed]
  23. 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]
  24. Li, Y.; Zhao, J.; Yu, S.; Wang, Z.; He, X.; Su, Y.; Guo, T.; Sheng, H.; Chen, J.; Zheng, Q.; et al. Extracellular Vesicles Long RNA Sequencing Reveals Abundant mRNA, circRNA, and lncRNA in Human Blood as Potential Biomarkers for Cancer Diagnosis. Clin. Chem. 2019, 65, 798–808. [Google Scholar] [CrossRef]
  25. 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]
  26. Guescini, M.; Genedani, S.; Stocchi, V.; Agnati, L.F. Astrocytes and Glioblastoma cells release exosomes carrying mtDNA. J. Neural Transm. 2009, 117, 1–4. [Google Scholar] [CrossRef]
  27. Belov, L.; Matic, K.J.; Hallal, S.; Best, O.G.; Mulligan, S.P.; Christopherson, R.I. Extensive surface protein profiles of extracellular vesicles from cancer cells may provide diagnostic signatures from blood samples. J. Extracell. Vesicles 2016, 5, 25355. [Google Scholar] [CrossRef]
  28. Boukouris, S.; Mathivanan, S. Exosomes in bodily fluids are a highly stable resource of disease biomarkers. Proteom.—Clin. Appl. 2015, 9, 358–367. [Google Scholar] [CrossRef]
  29. Li, S.; Li, Y.; Chen, B.; Zhao, J.; Yu, S.; Tang, Y.; Zheng, Q.; Li, Y.; Wang, P.; He, X.; et al. exoRBase: A database of circRNA, lncRNA and mRNA in human blood exosomes. Nucleic Acids Res. 2018, 46, D106–D112. [Google Scholar] [CrossRef] [PubMed]
  30. Liu, C.-J.; Xie, G.-Y.; Miao, Y.-R.; Xia, M.; Wang, Y.; Lei, Q.; Zhang, Q.; Guo, A.-Y. EVAtlas: A comprehensive database for ncRNA expression in human extracellular vesicles. Nucleic Acids Res. 2022, 50, D111–D117. [Google Scholar] [CrossRef] [PubMed]
  31. Liu, T.; Zhang, Q.; Zhang, J.; Li, C.; Miao, Y.R.; Lei, Q.; Li, Q.; Guo, A.Y. EVmiRNA: A database of miRNA profiling in extracellular vesicles. Nucleic Acids Res. 2019, 47, D89–D93. [Google Scholar] [CrossRef] [PubMed]
  32. Melling, G.E.; Carollo, E.; Conlon, R.; Simpson, J.C.; Carter, D.R.F. The Challenges and Possibilities of Extracellular Vesicles as Therapeutic Vehicles. Eur. J. Pharm. Biopharm. 2019, 144, 50–56. [Google Scholar] [CrossRef] [PubMed]
  33. Andreeva, O.E.; Shchegolev, Y.Y.; Scherbakov, A.M.; Mikhaevich, E.I.; Sorokin, D.V.; Gudkova, M.V.; Bure, I.V.; Kuznetsova, E.B.; Mikhaylenko, D.S.; Nemtsova, M.V.; et al. Secretion of Mutant DNA and mRNA by the Exosomes of Breast Cancer Cells. Molecules 2021, 26, 2499. [Google Scholar] [CrossRef] [PubMed]
  34. 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. 2017, 14, 305–316. [Google Scholar] [CrossRef]
  35. Liu, C.; Li, B.; Lin, H.; Yang, C.; Guo, J.; Cui, B.; Pan, W.; Feng, J.; Luo, T.; Chu, F.; et al. Multiplexed analysis of small extracellular vesicle-derived mRNAs by droplet digital PCR and machine learning improves breast cancer diagnosis. Biosens. Bioelectron. 2021, 194, 113615. [Google Scholar] [CrossRef]
  36. Abdulhussain, M.M.; Hasan, N.A.; Hussain, A.G. Interrelation of the Circulating and Tissue MicroRNA-21 with Tissue PDCD4 Expression and the Invasiveness of Iraqi Female Breast Tumors. Indian J. Clin. Biochem. 2017, 34, 26–38. [Google Scholar] [CrossRef]
  37. Ando, W.; Kikuchi, K.; Uematsu, T.; Yokomori, H.; Takaki, T.; Sogabe, M.; Kohgo, Y.; Otori, K.; Ishikawa, S.; Okazaki, I. Novel breast cancer screening: Combined expression of miR-21 and MMP-1 in urinary exosomes detects 95% of breast cancer without metastasis. Sci. Rep. 2019, 9, 13595. [Google Scholar] [CrossRef]
  38. 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]
  39. Liu, M.; Mo, F.; Song, X.; He, Y.; Yuan, Y.; Yan, J.; Yang, Y.; Huang, J.; Zhang, S. Exosomal hsa-miR-21-5p is a biomarker for breast cancer diagnosis. PeerJ 2021, 9, e12147. [Google Scholar] [CrossRef]
  40. Wang, J.; Zhang, Q.; Zhou, S.; Xu, H.; Wang, D.; Feng, J.; Zhao, J.; Zhong, S. Circular RNA expression in exosomes derived from breast cancer cells and patients. Epigenomics 2019, 11, 411–421. [Google Scholar] [CrossRef]
  41. Fiskaa, T.; Knutsen, E.; Nikolaisen, M.A.; Jørgensen, T.E.; Johansen, S.D.; Perander, M.; Seternes, O.M. Distinct Small RNA Signatures in Extracellular Vesicles Derived from Breast Cancer Cell Lines. PLoS ONE 2016, 11, e0161824. [Google Scholar] [CrossRef] [PubMed]
  42. 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, 2013, e201. [Google Scholar] [CrossRef] [PubMed]
  43. Feng, T.; Zhang, P.; Sun, Y.; Wang, Y.; Tong, J.; Dai, H.; Hua, Z. High throughput sequencing identifies breast cancer-secreted exosomal LncRNAs initiating pulmonary pre-metastatic niche formation. Gene 2019, 710, 258–264. [Google Scholar] [CrossRef]
  44. Ruhen, O.; Mirzai, B.; Clark, M.E.; Nguyen, B.; Salomon, C.; Erber, W.; Meehan, K. Comparison of Circulating Tumour DNA and Extracellular Vesicle DNA by Low-Pass Whole-Genome Sequencing Reveals Molecular Drivers of Disease in a Breast Cancer Patient. Biomedicines 2020, 9, 14. [Google Scholar] [CrossRef]
  45. De Carolis, S.; Storci, G.; Ceccarelli, C.; Savini, C.; Gallucci, L.; Sansone, P.; Santini, D.; Seracchioli, R.; Taffurelli, M.; Fabbri, F.; et al. HPV DNA Associates with Breast Cancer Malignancy and It Is Transferred to Breast Cancer Stromal Cells by Extracellular Vesicles. Front. Oncol. 2019, 9, 860. [Google Scholar] [CrossRef] [PubMed]
  46. Arroyo, J.D.; Chevillet, J.R.; Kroh, E.M.; Ruf, I.K.; Pritchard, C.C.; Gibson, D.F.; Mitchell, P.S.; Bennett, C.F.; Pogosova-Agadjanyan, E.L.; Stirewalt, D.L.; et al. Argonaute2 complexes carry a population of circulating microRNAs independent of vesicles in human plasma. Proc. Natl. Acad. Sci. USA 2011, 108, 5003–5008. [Google Scholar] [CrossRef]
  47. Crescitelli, R.; Lässer, C.; Szabó, T.G.; Kittel, A.; Eldh, M.; Dianzani, I.; Buzás, E.I.; Lötvall, J. Distinct RNA profiles in subpopulations of extracellular vesicles: Apoptotic bodies, microvesicles and exosomes. J. Extracell. Vesicles 2013, 2, 20677. [Google Scholar] [CrossRef]
  48. Uen, Y.; Wang, J.-W.; Wang, C.; Jhang, Y.; Chung, J.-Y.; Tseng, T.; Sheu, M.; Lee, S. Mining of potential microRNAs with clinical correlation—Regulation of syndecan-1 expression by miR-122-5p altered mobility of breast cancer cells and possible correlation with liver injury. Oncotarget 2018, 9, 28165–28175. [Google Scholar] [CrossRef]
  49. 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]
  50. Xun, J.; Du, L.; Gao, R.; Shen, L.; Wang, D.; Kang, L.; Chen, C.A.; 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]
  51. 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] [PubMed]
  52. Cereghetti, D.; Lee, P. Tumor-Derived Exosomes Contain microRNAs with Immunological Function: Implications for a Novel Immunosuppression Mechanism. MicroRNA 2014, 2, 194–204. [Google Scholar] [CrossRef] [PubMed]
  53. 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]
  54. Andreeva, O.E.; Sorokin, D.V.; Mikhaevich, E.I.; Bure, I.V.; Shchegolev, Y.Y.; Nemtsova, M.V.; Gudkova, M.V.; Scherbakov, A.M.; Krasil’nikov, M.A. Towards Unravelling the Role of ERα-Targeting miRNAs in the Exosome-Mediated Transferring of the Hormone Resistance. Molecules 2021, 26, 6661. [Google Scholar] [CrossRef]
  55. Lu, C.; Zhao, Y.; Wang, J.; Shi, W.; Dong, F.; Xin, Y.; Zhao, X.; Liu, C. Breast cancer cell-derived extracellular vesicles transfer miR-182-5p and promote breast carcinogenesis via the CMTM7/EGFR/AKT axis. Mol. Med. 2021, 27, 78. [Google Scholar] [CrossRef] [PubMed]
  56. Mihelich, B.L.; Dambal, S.; Lin, S.; Nonn, L. miR-182, of the miR-183 cluster family, is packaged in exosomes and is detected in human exosomes from serum, breast cells and prostate cells. Oncol. Lett. 2016, 12, 1197–1203. [Google Scholar] [CrossRef]
  57. 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. 2017, 44, 1741–1748. [Google Scholar] [CrossRef]
  58. Wang, M.; Zhang, H.; Yang, F.; Qiu, R.; Zhao, X.; Gong, Z.; Yu, W.; Zhou, B.; Shen, B.; Zhu, W. miR-188-5p suppresses cellular proliferation and migration via IL6ST: A potential noninvasive diagnostic biomarker for breast cancer. J. Cell. Physiol. 2020, 235, 4890–4901. [Google Scholar] [CrossRef]
  59. Wang, B.; Mao, J.H.; Wang, B.Y.; Wang, L.X.; Wen, H.Y.; Xu, L.J.; Fu, J.X.; Yang, H. Exosomal miR-1910-3p promotes proliferation, metastasis, and autophagy of breast cancer cells by targeting MTMR3 and activating the NF-κB signaling pathway. Cancer Lett. 2020, 489, 87–99. [Google Scholar] [CrossRef]
  60. Wu, K.; Feng, J.; Lyu, F.; Xing, F.; Sharma, S.; Liu, Y.; Wu, S.Y.; Zhao, D.; Tyagi, A.; Deshpande, R.P.; et al. Exosomal miR-19a and IBSP cooperate to induce osteolytic bone metastasis of estrogen receptor-positive breast cancer. Nat. Commun. 2021, 12, 5196. [Google Scholar] [CrossRef]
  61. Le, M.T.N.; Hamar, P.; Guo, C.; Basar, E.; Perdigão-Henriques, R.; Balaj, L.; Lieberman, J. miR-200–containing extracellular vesicles promote breast cancer cell metastasis. J. Clin. Investig. 2014, 124, 5109–5128. [Google Scholar] [CrossRef]
  62. Buschmann, D.; González, R.; Kirchner, B.; Mazzone, C.; Pfaffl, M.W.; Schelling, G.; Steinlein, O.; Reithmair, M. Glucocorticoid receptor overexpression slightly shifts microRNA expression patterns in triple-negative breast cancer. Int. J. Oncol. 2018, 52, 1765–1776. [Google Scholar] [CrossRef] [PubMed]
  63. Zhao, Y.; Jin, L.J.; Zhang, X.Y. Exosomal miRNA-205 promotes breast cancer chemoresistance and tumorigenesis through E2F1. Aging 2021, 13, 18498–18514. [Google Scholar] [CrossRef] [PubMed]
  64. Guo, L.; Zhu, Y.; Li, L.; Zhou, S.; Yin, G.; Yu, G.; Cui, H. Breast cancer cell-derived exosomal miR-20a-5p promotes the proliferation and differentiation of osteoclasts by targeting SRCIN1. Cancer Med. 2019, 8, 5687–5701. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, W.; Yuan, X.; Xu, A.; Zhu, X.; Zhan, Y.; Wang, S.; Liu, M. Human cancer cells suppress behaviors of endothelial progenitor cells through miR-21 targeting IL6R. Microvasc. Res. 2018, 120, 21–28. [Google Scholar] [CrossRef]
  66. Arisan, E.D.; Rencuzogullari, O.; Cieza-borrella, C.; Miralles Arenas, F.; Dwek, M.; Lange, S.; Uysal-onganer, P. MiR-21 is Required for the Epithelial–Mesenchymal Transition in MDA-MB-231 Breast Cancer Cells. Int. J. Mol. Sci. 2021, 22, 1557. [Google Scholar] [CrossRef]
  67. Lee, J.H.; Kim, J.A.; Kwon, M.H.; Kang, J.Y.; Rhee, W.J. In situ single step detection of exosome microRNA using molecular beacon. Biomaterials 2015, 54, 116–125. [Google Scholar] [CrossRef]
  68. 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]
  69. Guzman, N.; Agarwal, K.; Asthagiri, D.; Yu, L.; Saji, M.; Ringel, M.D.; Paulaitis, M.E. Breast Cancer–Specific miR Signature Unique to Extracellular Vesicles Includes “microRNA-like” tRNA Fragments. Mol. Cancer Res. 2015, 13, 891–901. [Google Scholar] [CrossRef]
  70. Bose, R.J.; Kumar, U.S.; Garcia-Marques, F.; Zeng, Y.; Habte, F.; McCarthy, J.R.; Pitteri, S.; Massoud, T.F.; Paulmurugan, R. Engineered Cell-Derived Vesicles Displaying Targeting Peptide and Functionalized with Nanocarriers for Therapeutic microRNA Delivery to Triple-Negative Breast Cancer in Mice. Adv. Healthc. Mater. 2021, 11, 2101387. [Google Scholar] [CrossRef]
  71. Wang, Z.; Zong, S.; Liu, Y.; Qian, Z.; Zhu, K.; Yang, Z.; Wang, Z.; Cui, Y. Simultaneous detection of multiple exosomal microRNAs for exosome screening based on rolling circle amplification. Nanotechnology 2020, 32, 085504. [Google Scholar] [CrossRef] [PubMed]
  72. Wang, H.; He, D.; Wan, K.; Sheng, X.; Cheng, H.; Huang, J.; Zhou, X.; He, X.; Wang, K. In situ multiplex detection of serum exosomal microRNAs using an all-in-one biosensor for breast cancer diagnosis. Analyst 2020, 145, 3289–3296. [Google Scholar] [CrossRef] [PubMed]
  73. Aslan, C.; Maralbashi, S.; Kahroba, H.; Asadi, M.; Soltani-Zangbar, M.S.; Javadian, M.; Shanehbandi, D.; Baradaran, B.; Darabi, M.; Kazemi, T. Docosahexaenoic acid (DHA) inhibits pro-angiogenic effects of breast cancer cells via down-regulating cellular and exosomal expression of angiogenic genes and microRNAs. Life Sci. 2020, 258, 118094. [Google Scholar] [CrossRef] [PubMed]
  74. 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]
  75. Das, K.; Paul, S.; Singh, A.; Ghosh, A.; Roy, A.; Ansari, S.A.; Prasad, R.; Mukherjee, A.; Sen, P. Triple-negative breast cancer-derived microvesicles transfer microRNA221 to the recipient cells and thereby promote epithelial-to-mesenchymal transition. J. Biol. Chem. 2019, 294, 13681–13696. [Google Scholar] [CrossRef]
  76. 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]
  77. Dan Yu, D.; Wu, Y.; Zhang, X.H.; Lv, M.M.; Chen, W.X.; Chen, X.; Yang, S.J.; Shen, H.; Zhong, S.L.; Tang, J.H.; et al. Exosomes from adriamycin-resistant breast cancer cells transmit drug resistance partly by delivering miR-222. Tumour Biol. 2016, 37, 3227–3235. [Google Scholar] [CrossRef]
  78. Chen, W.X.; Wang, D.D.; Zhu, B.; Zhu, Y.Z.; Zheng, L.; Feng, Z.Q.; Qin, X.H. Exosomal miR-222 from adriamycin-resistant MCF-7 breast cancer cells promote macrophages M2 polarization via PTEN/Akt to induce tumor progression. Aging 2021, 13, 10415–10430. [Google Scholar] [CrossRef]
  79. Du, J.; Fan, J.J.; Dong, C.; Li, H.T.; Ma, B.L. Inhibition effect of exosomes-mediated Let-7a on the development and metastasis of Triple negative breast cancer by down-regulating the expression of c-Myc. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 5301–5314. [Google Scholar] [CrossRef]
  80. Ding, J.; Xu, Z.; Zhang, Y.; Tan, C.; Hu, W.; Wang, M.; Xu, Y.; Tang, J. Exosome-mediated miR-222 transferring: An insight into NF-κB-mediated breast cancer metastasis. Exp. Cell Res. 2018, 369, 129–138. [Google Scholar] [CrossRef]
  81. Morad, G.; Daisy, C.C.; Otu, H.H.; Libermann, T.A.; Dillon, S.T.; Moses, M.A. Cdc42-Dependent Transfer of mir301 from Breast Cancer-Derived Extracellular Vesicles Regulates the Matrix Modulating Ability of Astrocytes at the Blood–Brain Barrier. Int. J. Mol. Sci. 2020, 21, 3851. [Google Scholar] [CrossRef] [PubMed]
  82. 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] [PubMed]
  83. 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]
  84. 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]
  85. Moloney, B.M.; Gilligan, K.E.; Joyce, D.P.; O’Neill, C.P.; O’Brien, K.P.; Khan, S.; Glynn, C.L.; Waldron, R.M.; Maguire, C.M.; Holian, E.; et al. Investigating the Potential and Pitfalls of EV-Encapsulated MicroRNAs as Circulating Biomarkers of Breast Cancer. Cells 2020, 9, 141. [Google Scholar] [CrossRef]
  86. Xin, Y.; Wang, X.; Meng, K.; Ni, C.; Lv, Z.; Guan, D. Identification of exosomal miR-455-5p and miR-1255a as therapeutic targets for breast cancer. Biosci. Rep. 2020, 40, 20190303. [Google Scholar] [CrossRef]
  87. 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]
  88. Li, Y.; Liang, Y.; Sang, Y.; Song, X.; Zhang, H.; Liu, Y.; Jiang, L.; Yang, Q. MiR-770 suppresses the chemo-resistance and metastasis of triple negative breast cancer via direct targeting of STMN1. Cell Death Dis. 2018, 9, 14. [Google Scholar] [CrossRef]
  89. 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]
  90. Hannafon, B.N.; Carpenter, K.J.; Berry, W.L.; Janknecht, R.; Dooley, W.C.; Ding, W.Q. Exosome-mediated microRNA signaling from breast cancer cells is altered by the anti-angiogenesis agent docosahexaenoic acid (DHA). Mol. Cancer 2015, 14, 133. [Google Scholar] [CrossRef]
  91. Kia, V.; Paryan, M.; Mortazavi, Y.; Biglari, A.; Mohammadi-Yeganeh, S. Evaluation of exosomal miR-9 and miR-155 targeting PTEN and DUSP14 in highly metastatic breast cancer and their effect on low metastatic cells. J. Cell. Biochem. 2019, 120, 5666–5676. [Google Scholar] [CrossRef] [PubMed]
  92. 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] [PubMed]
  93. 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]
  94. Zhang, W.; Liu, H.; Jiang, J.; Yang, Y.; Wang, W.; Jia, Z. CircRNA circFOXK2 facilitates oncogenesis in breast cancer via IGF2BP3/miR-370 axis. Aging 2021, 13, 18978. [Google Scholar] [CrossRef] [PubMed]
  95. Li, M.; Zou, X.; Xia, T.; Wang, T.; Liu, P.; Zhou, X.; Wang, S.; Zhu, W. A five-miRNA panel in plasma was identified for breast cancer diagnosis. Cancer Med. 2019, 8, 7006–7017. [Google Scholar] [CrossRef]
  96. Chen, Y.; Zhai, L.-Y.; Zhang, L.-M.; Ma, X.-S.; Liu, Z.; Li, M.-M.; Chen, J.-X.; Duan, W.-J. Breast cancer plasma biopsy by in situ determination of exosomal microRNA-1246 with a molecular beacon. Analyst 2021, 146, 2264–2276. [Google Scholar] [CrossRef]
  97. Zhai, L.-Y.; Li, M.-X.; Pan, W.-L.; Chen, Y.; Li, M.-M.; Pang, J.-X.; Zheng, L.; Chen, J.-X.; Duan, W.-J. In Situ Detection of Plasma Exosomal MicroRNA-1246 for Breast Cancer Diagnostics by a Au Nanoflare Probe. ACS Appl. Mater. Interfaces 2018, 10, 39478–39486. [Google Scholar] [CrossRef]
  98. 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]
  99. Ni, Q.; Stevic, I.; Pan, C.; Müller, V.; Oliviera-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]
  100. Ahmed, S.H.; Espinoza-Sánchez, N.A.; El-Damen, A.; Fahim, S.A.; Badawy, M.A.; Greve, B.; El-Shinawi, M.; Götte, M.; Ibrahim, S.A. Small extracellular vesicle-encapsulated miR-181b-5p, miR-222-3p and let-7a-5p: Next generation plasma biopsy-based diagnostic biomarkers for inflammatory breast cancer. PLoS ONE 2021, 16, e0250642. [Google Scholar] [CrossRef]
  101. 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]
  102. 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, 2019, 21. [Google Scholar] [CrossRef] [PubMed]
  103. 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]
  104. 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]
  105. Zou, X.; Li, M.; Huang, Z.; Zhou, X.; Liu, Q.; Xia, T.; Zhu, W. Circulating miR-532-502 cluster derived from chromosome X as biomarkers for diagnosis of breast cancer. Gene 2020, 722, 144104. [Google Scholar] [CrossRef]
  106. Kim, M.W.; Park, S.; Lee, H.; Gwak, H.; Hyun, K.A.; Kim, J.Y.; Jung, H.I.; II Kim, S. Multi-miRNA panel of tumor-derived extracellular vesicles as promising diagnostic biomarkers of early-stage breast cancer. Cancer Sci. 2021, 112, 5078–5087. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, X.; Qian, T.; Bao, S.; Zhao, H.; Chen, H.; Xing, Z.; Li, Y.; Zhang, M.; Meng, X.; Wang, C.; et al. Circulating exosomal miR-363-5p inhibits lymph node metastasis by downregulating PDGFB and serves as a potential noninvasive biomarker for breast cancer. Mol. Oncol. 2021, 15, 2466–2479. [Google Scholar] [CrossRef]
  108. 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]
  109. Li, D.; Wang, J.; Ma, L.J.; Yang, H.B.; Jing, J.F.; Jia, M.M.; Zhang, X.J.; Guo, F.; Gao, J.N. Identification of serum exosomal miR-148a as a novel prognostic biomarker for breast cancer. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 7303–7309. [Google Scholar] [CrossRef]
  110. Yuan, X.; Qian, N.; Ling, S.; Li, Y.; Sun, W.; Li, J.; Du, R.; Zhong, G.; Liu, C.; Yu, G.; et al. Breast cancer exosomes contribute to pre-metastatic niche formation and promote bone metastasis of tumor cells. Theranostics 2021, 11, 1429–1445. [Google Scholar] [CrossRef]
  111. Li, S.; Zhang, M.; Xu, F.; Wang, Y.; Leng, D. Detection significance of miR-3662, miR-146a, and miR-1290 in serum exosomes of breast cancer patients. J. Cancer Res. Ther. 2021, 17, 749–755. [Google Scholar] [CrossRef] [PubMed]
  112. 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] [PubMed]
  113. Eichelser, C.; Stückrath, I.; Müller, V.; Milde-Langosch, K.; Wikman, H.; Pantel, K.; Schwarzenbach, H.; Eichelser, C.; Stückrath, I.; Müller, V.; et al. Increased serum levels of circulating exosomal microRNA-373 in receptor-negative breast cancer patients. Oncotarget 2014, 5, 9650–9663. [Google Scholar] [CrossRef]
  114. Yang, Q.; Zhao, S.; Shi, Z.; Cao, L.; Liu, J.; Pan, T.; Zhou, D.; Zhang, J. Chemotherapy-elicited exosomal miR-378a-3p and miR-378d promote breast cancer stemness and chemoresistance via the activation of EZH2/STAT3 signaling. J. Exp. Clin. Cancer Res. 2021, 40, 120. [Google Scholar] [CrossRef] [PubMed]
  115. Bao, S.; Hu, T.; Liu, J.; Su, J.; Sun, J.; Ming, Y.; Li, J.; Wu, N.; Chen, H.; Zhou, M. Genomic instability-derived plasma extracellular vesicle-microRNA signature as a minimally invasive predictor of risk and unfavorable prognosis in breast cancer. J. Nanobiotechnol. 2021, 19, 22. [Google Scholar] [CrossRef] [PubMed]
  116. Curtaz, C.J.; Reifschläger, L.; Strähle, L.; Feldheim, J.; Feldheim, J.J.; Schmitt, C.; Kiesel, M.; Herbert, S.-L.; Wöckel, A.; Meybohm, P.; et al. Analysis of microRNAs in Exosomes of Breast Cancer Patients in Search of Molecular Prognostic Factors in Brain Metastases. Int. J. Mol. Sci. 2022, 23, 3683. [Google Scholar] [CrossRef]
  117. 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]
  118. Inubushi, S.; Kawaguchi, H.; Mizumoto, S.; Kunihisa, T.; Baba, M.; Kitayama, Y.; Takeuchi, T.; Hoffman, R.M.; Sasaki, R. Oncogenic miRNAs Identified in Tear Exosomes from Metastatic Breast Cancer Patients. Anticancer Res. 2020, 40, 3091–3096. [Google Scholar] [CrossRef]
  119. Hirschfeld, M.; Rücker, G.; Weiß, D.; Berner, K.; Ritter, A.; Jäger, M.; Erbes, T. Urinary Exosomal MicroRNAs as Potential Non-invasive Biomarkers in Breast Cancer Detection. Mol. Diagn. Ther. 2020, 24, 215–232. [Google Scholar] [CrossRef]
  120. Koi, Y.; Tsutani, Y.; Nishiyama, Y.; Ueda, D.; Ibuki, Y.; Sasada, S.; Akita, T.; Masumoto, N.; Kadoya, T.; Yamamoto, Y.; et al. Predicting the presence of breast cancer using circulating small RNAs, including those in the extracellular vesicles. Cancer Sci. 2020, 111, 2104–2115. [Google Scholar] [CrossRef]
  121. Galli De Amorim, M.; Branco, G.; Valieris, R.; Tarcitano, E.; Tojal Da Silva, I.; Ferreira De Araújo, L.; Noronha Nunes, D.; Dias-Neto, E. The impact of HER2 overexpression on the miRNA and circRNA transcriptomes in two breast cell lines and their vesicles. Pharmacogenomics 2019, 20, 493–502. [Google Scholar] [CrossRef] [PubMed]
  122. 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]
  123. Singh, R.; Pochampally, R.; Watabe, K.; Lu, Z.; Mo, Y.-Y. Exosome-mediated transfer of miR-10b promotes cell invasion in breast cancer. Mol. Cancer 2014, 13, 256. [Google Scholar] [CrossRef] [PubMed]
  124. Dai, X.; Cheng, H.; Bai, Z.; Li, J. Breast Cancer Cell Line Classification and Its Relevance with Breast Tumor Subtyping. J. Cancer 2017, 8, 3131–3141. [Google Scholar] [CrossRef]
  125. Petrović, N. miR-21 Might be Involved in Breast Cancer Promotion and Invasion Rather than in Initial Events of Breast Cancer Development. Mol. Diagn. Ther. 2016, 20, 97–110. [Google Scholar] [CrossRef]
  126. Yan, L.-X.; Huang, X.-F.; Shao, Q.; Huang, M.-Y.; Deng, L.; Wu, Q.-L.; Zeng, Y.-X.; Shao, J.-Y. MicroRNA miR-21 overexpression in human breast cancer is associated with advanced clinical stage, lymph node metastasis and patient poor prognosis. RNA 2008, 14, 2348–2360. [Google Scholar] [CrossRef]
  127. 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]
  128. Shah, S.H.; Miller, P.; Garcia-Contreras, M.; Ao, Z.; Machlin, L.; Issa, E.; El-Ashry, D. Hierarchical paracrine interaction of breast cancer associated fibroblasts with cancer cells via hMAPK-microRNAs to drive ER-negative breast cancer phenotype. Cancer Biol. Ther. 2015, 16, 1671–1681. [Google Scholar] [CrossRef]
  129. Palviainen, M.; Saraswat, M.; Varga, Z.; Kitka, D.; Neuvonen, M.; Puhka, M.; Joenväärä, S.; Renkonen, R.; Nieuwland, R.; Takatalo, M.; et al. Extracellular vesicles from human plasma and serum are carriers of extravesicular cargo—Implications for biomarker discovery. PLoS ONE 2020, 15, e0236439. [Google Scholar] [CrossRef]
  130. Sun, Y.; Saito, K.; Saito, Y. Lipid Profile Characterization and Lipoprotein Comparison of Extracellular Vesicles from Human Plasma and Serum. Metabolites 2019, 9, 259. [Google Scholar] [CrossRef]
  131. 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]
  132. Wu, H.; Wang, Q.; Zhong, H.; Li, L.; Zhang, Q.; Huang, Q.; Yu, Z. Differentially expressed microRNAs in exosomes of patients with breast cancer revealed by next-generation sequencing. Oncol. Rep. 2020, 43, 240–250. [Google Scholar] [CrossRef] [PubMed]
  133. 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]
  134. Li, Q.; Eades, G.; Yao, Y.; Zhang, Y.; Zhou, Q. Characterization of a Stem-like Subpopulation in Basal-like Ductal Carcinoma in Situ (DCIS) Lesions *. J. Biol. Chem. 2014, 289, 1303–1312. [Google Scholar] [CrossRef] [PubMed]
  135. Jc Bose, R.; Uday Kumar, S.; Zeng, Y.; Afjei, R.; Robinson, E.; Lau, K.; Bermudez, A.; Habte, F.; Pitteri, S.J.; Sinclair, R.; et al. Tumor Cell-Derived Extracellular Vesicle-Coated Nanocarriers: An Efficient Theranostic Platform for the Cancer-Specific Delivery of Anti-miR-21 and Imaging Agents. ACS Nano 2018, 12, 10817–10832. [Google Scholar] [CrossRef] [PubMed]
  136. Taghikhani, A.; Hassan, Z.M.; Ebrahimi, M.; Moazzeni, S.M. microRNA modified tumor-derived exosomes as novel tools for maturation of dendritic cells. J. Cell. Physiol. 2019, 234, 9417–9427. [Google Scholar] [CrossRef]
  137. Zhao, L.; Gu, C.; Gan, Y.; Shao, L.; Chen, H.; Zhu, H. Exosome-mediated siRNA delivery to suppress postoperative breast cancer metastasis. J. Control. Release 2020, 318, 1–15. [Google Scholar] [CrossRef]
  138. Lobos-González, L.; Bustos, R.; Campos, A.; Silva, V.; Silva, V.; Jeldes, E.; Salomon, C.; Varas-Godoy, M.; Cáceres-Verschae, A.; Duran, E.; et al. Exosomes released upon mitochondrial ASncmtRNA knockdown reduce tumorigenic properties of malignant breast cancer cells. Sci. Rep. 2020, 10, 343. [Google Scholar] [CrossRef]
  139. Xu, X.; Su, B.; Xie, C.; Wei, S.; Zhou, Y.; Liu, H.; Dai, W.; Cheng, P.; Wang, F.; Xu, X.; et al. Sonic Hedgehog-Gli1 Signaling Pathway Regulates the Epithelial Mesenchymal Transition (EMT) by Mediating a New Target Gene, S100A4, in Pancreatic Cancer Cells. PLoS ONE 2014, 9, e96441. [Google Scholar] [CrossRef]
  140. Zhou, Y.; Yamamoto, Y.; Takeshita, F.; Yamamoto, T.; Xiao, Z.; Ochiya, T. Delivery of miR-424-5p via Extracellular Vesicles Promotes the Apoptosis of MDA-MB-231 TNBC Cells in the Tumor Microenvironment. Int. J. Mol. Sci. 2021, 22, 844. [Google Scholar] [CrossRef]
  141. 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, 1–14. [Google Scholar] [CrossRef]
  142. Xia, W.; Liu, Y.; Cheng, T.; Xu, T.; Dong, M.; Hu, X. Extracellular Vesicles Carry lncRNA SNHG16 to Promote Metastasis of Breast Cancer Cells via the miR-892b/PPAPDC1A Axis. Front. Cell Dev. Biol. 2021, 9, 1319. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, Y.L.; Liu, L.C.; Hung, Y.; Chen, C.J.; Lin, Y.Z.; Wu, W.R.; Wang, S.C. Long non-coding RNA HOTAIR in circulatory exosomes is correlated with ErbB2/HER2 positivity in breast cancer. Breast 2019, 46, 64–69. [Google Scholar] [CrossRef] [PubMed]
  144. 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]
  145. 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. 2018, 53, 1013–1026. [Google Scholar] [CrossRef]
  146. Han, M.; Gu, Y.; Lu, P.; Li, J.; Cao, H.; Li, X.; Qian, X.; Yu, C.; Yang, Y.; Yang, X.; et al. Exosome-mediated lncRNA AFAP1-AS1 promotes trastuzumab resistance through binding with AUF1 and activating ERBB2 translation. Mol. Cancer 2020, 19, 1–18. [Google Scholar] [CrossRef]
  147. 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. 2021, 11, 1962. [Google Scholar] [CrossRef]
  148. 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. 2019, 25, 2211–2220. [Google Scholar] [CrossRef]
  149. 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]
  150. 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. Sci. 2016, 20, 4362–4368. [Google Scholar]
  151. Lu, Y.; Chen, L.; Li, L.; Cao, Y. Exosomes Derived from Brain Metastatic Breast Cancer Cells Destroy the Blood-Brain Barrier by Carrying lncRNA GS1-600G8.5. Biomed. Res. Int. 2020, 2020, 7461727. [Google Scholar] [CrossRef] [PubMed]
  152. Zhou, D.P.; Gu, J.; Wang, Y.P.; Wu, H.G.; Cheng, W.; Wang, Q.P.; Zheng, G.P.; Wang, X.D. Long non-coding RNA NEAT1 transported by extracellular vesicles contributes to breast cancer development by sponging microRNA-141-3p and regulating KLF12. Cell Biosci. 2021, 11, 68. [Google Scholar] [CrossRef] [PubMed]
  153. Chen, F.; Chen, J.; Yang, L.; Liu, J.; Zhang, X.; Zhang, Y.; Tu, Q.; Yin, D.; Lin, D.; Wong, P.P.; et al. Extracellular vesicle-packaged HIF-1α-stabilizing lncRNA from tumour-associated macrophages regulates aerobic glycolysis of breast cancer cells. Nat. Cell Biol. 2019, 21, 498–510. [Google Scholar] [CrossRef] [PubMed]
  154. 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. Biotechnol. 2020, 191, 1084–1099. [Google Scholar] [CrossRef] [PubMed]
  155. Xing, F.; Liu, Y.; Wu, S.Y.; Wu, K.; Sharma, S.; Mo, Y.Y.; Feng, J.; Sanders, S.; Jin, G.; Singh, R.; et al. Loss of XIST in Breast Cancer Activates MSN-c-Met and Reprograms Microglia via Exosomal miRNA to Promote Brain Metastasis. Cancer Res. 2018, 78, 4316–4330. [Google Scholar] [CrossRef]
  156. 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]
  157. Chen, T.; Wang, X.; Li, C.; Zhang, H.; Liu, Y.; Han, D.; Li, Y.; Li, Z.; Luo, D.; Zhang, N.; et al. CircHIF1A regulated by FUS accelerates triple-negative breast cancer progression by modulating NFIB expression and translocation. Oncogene 2021, 40, 2756–2771. [Google Scholar] [CrossRef]
  158. 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]
  159. Yang, S.J.; Wang, D.D.; Zhou, S.Y.; Zhang, Q.; Wang, J.Y.; Zhong, S.L.; Da Zhang, H.; Wang, X.Y.; Xia, X.; Chen, W.; et al. Identification of circRNA–miRNA networks for exploring an underlying prognosis strategy for breast cancer. Epigenomics 2020, 12, 101–125. [Google Scholar] [CrossRef]
  160. 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]
  161. Keup, C.; Suryaprakash, V.; Storbeck, M.; Hoffmann, O.; Kimmig, R.; Kasimir-bauer, S. Longitudinal Multi-Parametric Liquid Biopsy Approach Identifies Unique Features of Circulating Tumor Cell, Extracellular Vesicle, and Cell-Free DNA Characterization for Disease Monitoring in Metastatic Breast Cancer Patients. Cells 2021, 10, 212. [Google Scholar] [CrossRef] [PubMed]
  162. Hu, M.; Brown, V.; Jackson, J.M.; Wijerathne, H.; Pathak, H.; Koestler, D.C.; Nissen, E.; Hupert, M.L.; Muller, R.; Godwin, A.K.; et al. Assessing Breast Cancer Molecular Subtypes Using Extracellular Vesicles’ mRNA. Anal. Chem. 2023, 95, 7665–7675. [Google Scholar] [CrossRef] [PubMed]
  163. Yao, B.; Qu, S.; Hu, R.; Gao, W.; Jin, S.; Ju, J.; Zhao, Q. Delivery of platelet TPM3 mRNA into breast cancer cells via microvesicles enhances metastasis. FEBS Open Biol. 2019, 9, 2159–2169. [Google Scholar] [CrossRef] [PubMed]
  164. Gorji-Bahri, G.; Moradtabrizi, N.; Vakhshiteh, F.; Hashemi, A. Validation of common reference genes stability in exosomal mRNA-isolated from liver and breast cancer cell lines. Cell Biol. Int. 2021, 45, 1098–1110. [Google Scholar] [CrossRef]
  165. Rabas, N.; Palmer, S.; Mitchell, L.; Ismail, S.; Gohlke, A.; Riley, J.S.; Tait, S.W.G.; Gammage, P.; Soares, L.L.; Macpherson, I.R.; et al. Pink1 drives production of mtDNA-containing extracellular vesicles to promote invasiveness. J. Cell Biol. 2021, 220, e202006049. [Google Scholar] [CrossRef]
  166. Abad, E.; Lyakhovich, A. Movement of Mitochondria with Mutant DNA through Extracellular Vesicles Helps Cancer Cells Acquire Chemoresistance. ChemMedChem 2021, 17, e202100642. [Google Scholar] [CrossRef]
  167. 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]
  168. Wang, J.H.; Forterre, A.V.; Zhao, J.; Frimannsson, D.O.; Delcayre, A.; Antes, T.J.; Efron, B.; Jeffrey, S.S.; Pegram, M.D.; Matin, A.C. Anti-HER2 scFv-Directed Extracellular Vesicle-Mediated mRNA-Based Gene Delivery Inhibits Growth of HER2-Positive Human Breast Tumor Xenografts by Prodrug Activation. Mol. Cancer Ther. 2018, 17, 1133–1142. [Google Scholar] [CrossRef]
  169. Hurwitz, S.N.; Rider, M.A.; Bundy, J.L.; Liu, X.; Singh, R.K.; Meckes, D.G., Jr. Proteomic profiling of NCI-60 extracellular vesicles uncovers common protein cargo and cancer type-specific biomarkers. Oncotarget 2016, 7, 86999–87015. [Google Scholar] [CrossRef]
  170. Staubach, S.; Razawi, H.; Hanisch, F.G. Proteomics of MUC1-containing lipid rafts from plasma membranes and exosomes of human breast carcinoma cells MCF-7. Proteomics 2009, 9, 2820–2835. [Google Scholar] [CrossRef]
  171. 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] [PubMed]
  172. Antonyak, M.A.; Li, B.; Boroughs, L.K.; Johnson, J.L.; Druso, J.E.; Bryant, K.L.; Holowka, D.A.; Cerione, R.A. Cancer cell-derived microvesicles induce transformation by transferring tissue transglutaminase and fibronectin to recipient cells. Proc. Natl. Acad. Sci. USA 2011, 108, 4852–4857. [Google Scholar] [CrossRef]
  173. Tutanov, O.; Orlova, E.; Proskura, K.; Grigor’eva, A.; Yunusova, N.; Tsentalovich, Y.; Alexandrova, A.; Tamkovich, S. Proteomic Analysis of Blood Exosomes from Healthy Females and Breast Cancer Patients Reveals an Association between Different Exosomal Bioactivity on Non-tumorigenic Epithelial Cell and Breast Cancer Cell Migration in Vitro. Biomolecules 2020, 10, 495. [Google Scholar] [CrossRef] [PubMed]
  174. Vinik, Y.; Ortega, F.G.; Mills, G.B.; Lu, Y.; Jurkowicz, M.; Halperin, S.; Aharoni, M.; Gutman, M.; Lev, S. Proteomic analysis of circulating extracellular vesicles identifies potential markers of breast cancer progression, recurrence, and response. Sci. Adv. 2020, 6, eaba5714. [Google Scholar] [CrossRef] [PubMed]
  175. 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] [PubMed]
  176. Rontogianni, S.; Synadaki, E.; Li, B.; Liefaard, M.C.; Lips, E.H.; Wesseling, J.; Wu, W.; Altelaar, M. Proteomic profiling of extracellular vesicles allows for human breast cancer subtyping. Commun. Biol. 2019, 2, 325. [Google Scholar] [CrossRef]
  177. 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, 2173. [Google Scholar] [CrossRef]
  178. Martínez-Greene, J.A.; Hernández-Ortega, K.; Quiroz-Baez, R.; Resendis-Antonio, O.; Pichardo-Casas, I.; Sinclair, D.A.; Budnik, B.; Hidalgo-Miranda, A.; Uribe-Querol, E.; del Pilar Ramos-Godínez, M.; et al. Quantitative proteomic analysis of extracellular vesicle subgroups isolated by an optimized method combining polymer-based precipitation and size exclusion chromatography. J. Extracell. Vesicles 2021, 10, e12087. [Google Scholar] [CrossRef]
  179. Dalla, P.V.; Santos, J.; Milthorpe, B.K.; Padula, M.P. Selectively-Packaged Proteins in Breast Cancer Extracellular Vesicles Involved in Metastasis. Int. J. Mol. Sci. 2020, 21, 4990. [Google Scholar] [CrossRef]
  180. Vardaki, I.; Ceder, S.; Rutishauser, D.; Baltatzis, G.; Foukakis, T.; Panaretakis, T.; Vardaki, I.; Ceder, S.; Rutishauser, D.; Baltatzis, G.; et al. Periostin is identified as a putative metastatic marker in breast cancer-derived exosomes. Oncotarget 2016, 7, 74966–74978. [Google Scholar] [CrossRef]
  181. 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] [PubMed]
  182. Palazzolo, G.; Albanese, N.N.; Di Cara, G.; Gygax, D.; Vittorelli, M.L.; Pucci-Minafra, I. Proteomic analysis of exosome-like vesicles derived from breast cancer cells. Anticancer Res. 2012, 32, 847–860. [Google Scholar] [PubMed]
  183. Heinemann, M.L.; Ilmer, M.; Silva, L.P.; Hawke, D.H.; Recio, A.; Vorontsova, M.A.; Alt, E.; Vykoukal, J. Benchtop isolation and characterization of functional exosomes by sequential filtration. J. Chromatogr. A 2014, 1371, 125–135. [Google Scholar] [CrossRef]
  184. Clark, D.J.; Fondrie, W.E.; Liao, Z.; Hanson, P.I.; Fulton, A.; Mao, L.; Yang, A.J. Redefining the Breast Cancer Exosome Proteome by Tandem Mass Tag Quantitative Proteomics and Multivariate Cluster Analysis. Anal. Chem. 2015, 87, 10462–10469. [Google Scholar] [CrossRef] [PubMed]
  185. 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-Biomembr. 2014, 1838, 2954–2965. [Google Scholar] [CrossRef]
  186. Griffiths, S.G.; Cormier, M.T.; Clayton, A.; Doucette, A.A. Differential Proteome Analysis of Extracellular Vesicles from Breast Cancer Cell Lines by Chaperone Affinity Enrichment. Proteomes 2017, 5, 25. [Google Scholar] [CrossRef]
  187. Koh, Y.Q.; Ng, D.Q.; Ng, C.C.; Boey, A.; Wei, M.; Sze, S.K.; Ho, H.K.; Acharya, M.; Limoli, C.L.; Chan, A. Extracellular Vesicle Proteome of Breast Cancer Patients with and without Cognitive Impairment Following Anthracycline-based Chemotherapy: An Exploratory Study. Biomark. Insights 2021, 16, 11772719211018204. [Google Scholar] [CrossRef]
  188. Kruger, S.; Elmageed, Z.Y.A.; Hawke, D.H.; Wörner, P.M.; Jansen, D.A.; Abdel-Mageed, A.B.; Alt, E.U.; Izadpanah, R. Molecular characterization of exosome-like vesicles from breast cancer cells. BMC Cancer 2014, 14, 1–10. [Google Scholar] [CrossRef]
  189. Tryndyak, V.P.; Kovalchuk, O.; Pogribny, I.P. Loss of DNA methylation and histone H4 lysine 20 trimethylation in human breast cancer cells is associated with aberrant expression of DNA methyltransferase 1, Suv4-20h2 histone methyltransferase and methyl-binding proteins. Cancer Biol. Ther. 2006, 5, 65–70. [Google Scholar] [CrossRef]
  190. Coticchia, C.M.; Revankar, C.M.; Deb, T.B.; Dickson, R.B.; Johnson, M.D. Calmodulin modulates Akt activity in human breast cancer cell lines. Breast Cancer Res. Treat. 2008, 115, 545–560. [Google Scholar] [CrossRef]
  191. Skotland, T.; Sagini, K.; Sandvig, K.; Llorente, A. An emerging focus on lipids in extracellular vesicles. Adv. Drug Deliv. Rev. 2020, 159, 308–321. [Google Scholar] [CrossRef] [PubMed]
  192. Skotland, T.; Hessvik, N.P.; Sandvig, K.; Llorente, A. Exosomal lipid composition and the role of ether lipids and phosphoinositides in exosome biology. J. Lipid Res. 2019, 60, 9–18. [Google Scholar] [CrossRef] [PubMed]
  193. Roberg-Larsen, H.; Lund, K.; Seterdal, K.E.; Solheim, S.; Vehus, T.; Solberg, N.; Krauss, S.; Lundanes, E.; Wilson, S.R. Mass spectrometric detection of 27-hydroxycholesterol in breast cancer exosomes. J. Steroid Biochem. Mol. Biol. 2017, 169, 22–28. [Google Scholar] [CrossRef] [PubMed]
  194. Nishida-Aoki, N.; Izumi, Y.; Takeda, H.; Takahashi, M.; Ochiya, T.; Bamba, T. Lipidomic Analysis of Cells and Extracellular Vesicles from High- and Low-Metastatic Triple-Negative Breast Cancer. Metabolites 2020, 10, 67. [Google Scholar] [CrossRef]
  195. Góss Santos, T.; Regina Martins, V.; Noeli Maroso Hajj, G.; Di Sansebastiano, G.-P.; Gaballo, A. Unconventional Secretion of Heat Shock Proteins in Cancer. Int. J. Mol. Sci. 2017, 18, 946. [Google Scholar] [CrossRef]
  196. Zhang, D.H.; Tai, L.K.; Wong, L.L.; Sethi, S.K.; Koay, E.S.C. Proteomics of breast cancer: Enhanced expression of cytokeratin19 in human epidermal growth factor receptor type 2 positive breast tumors. Proteomics 2005, 5, 1797–1805. [Google Scholar] [CrossRef]
  197. Olofsson, M.H.; Ueno, T.; Pan, Y.; Xu, R.; Cai, F.; Van Der Kuip, H.; Muerdter, T.E.; Sonnenberg, M.; Aulitzky, W.E.; Schwarz, S.; et al. Cytokeratin-18 Is a Useful Serum Biomarker for Early Determination of Response of Breast Carcinomas to Chemotherapy. Clin. Cancer Res. 2007, 13, 3198–3206. [Google Scholar] [CrossRef]
  198. Muñoz-Pinedo, C.; El Mjiyad, N.; Ricci, J.E. Cancer metabolism: Current perspectives and future directions. Cell Death Dis. 2012, 3, e248. [Google Scholar] [CrossRef]
  199. Becker, A.; Thakur, B.K.; Weiss, J.M.; Kim, H.S.; Peinado, H.; Lyden, D. Extracellular Vesicles in Cancer: Cell-to-Cell Mediators of Metastasis. Cancer Cell 2016, 30, 836–848. [Google Scholar] [CrossRef]
  200. Tian, F.; Zhang, S.; Liu, C.; Han, Z.; Liu, Y.; Deng, J.; Li, Y.; Wu, X.; Cai, L.; Qin, L.; et al. Protein analysis of extracellular vesicles to monitor and predict therapeutic response in metastatic breast cancer. Nat. Commun. 2021, 12, 2536. [Google Scholar] [CrossRef]
  201. Tkach, M.; Théry, C. Communication by Extracellular Vesicles: Where We Are and Where We Need to Go. Cell 2016, 164, 1226–1232. [Google Scholar] [CrossRef] [PubMed]
  202. Horiuchi, K.; Amizuka, N.; Takeshita, S.; Takamatsu, H.; Katsuura, M.; Ozawa, H.; Toyama, Y.; Bonewald, L.F.; Kudo, A. Identification and Characterization of a Novel Protein, Periostin, with Restricted Expression to Periosteum and Periodontal Ligament and Increased Expression by Transforming Growth Factor β. J. Bone Miner. Res. 1999, 14, 1239–1249. [Google Scholar] [CrossRef] [PubMed]
  203. Shao, R.; Bao, S.; Bai, X.; Blanchette, C.; Anderson, R.M.; Dang, T.; Gishizky, M.L.; Marks, J.R.; Wang, X.-F. Acquired Expression of Periostin by Human Breast Cancers Promotes Tumor Angiogenesis through Up-Regulation of Vascular Endothelial Growth Factor Receptor 2 Expression. Mol. Cell. Biol. 2004, 24, 3992–4003. [Google Scholar] [CrossRef]
  204. 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] [PubMed]
  205. Palviainen, M.; Saari, H.; Kärkkäinen, O.; Pekkinen, J.; Auriola, S.; Yliperttula, M.; Puhka, M.; Hanhineva, K.; Siljander, P.R.-M. Metabolic signature of extracellular vesicles depends on the cell culture conditions. J. Extracell. Vesicles 2019, 8, 1596669. [Google Scholar] [CrossRef] [PubMed]
  206. 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] [PubMed]
  207. Ilver, D.; Johansson, P.; Miller-Podraza, H.; Nyholm, P.G.; Teneberg, S.; Karlsson, K.A. Bacterium–Host Protein–Carbohydrate Interactions. Methods Enzymol. 2003, 363, 134–157. [Google Scholar] [CrossRef]
  208. Pinho, S.S.; Reis, C.A. Glycosylation in cancer: Mechanisms and clinical implications. Nat. Rev. Cancer 2015, 15, 540–555. [Google Scholar] [CrossRef]
  209. Marcos, N.T.; Bennett, E.P.; Gomes, J.; Magalhaes, A.; Gomes, C.; David, L.; Dar, I.; Jeanneau, C.; DeFrees, S.; Krustrup, D.; et al. ST6GalNAc-I controls expression of sialyl-Tn antigen in gastrointestinal tissues. Front. Biosci.-Elit. 2011, 3, 1443–1455. [Google Scholar] [CrossRef]
  210. Wang, Y.; Ju, T.; Ding, X.; Xia, B.; Wang, W.; Xia, L.; He, M.; Cummings, R.D. Cosmc is an essential chaperone for correct protein O-glycosylation. Proc. Natl. Acad. Sci. USA 2010, 107, 9228–9233. [Google Scholar] [CrossRef]
  211. Kumamoto, K.; Goto, Y.; Sekikawa, K.; Takenoshita, S.; Ishida, N.; Kawakita, M.; Kannagi, R. Increased Expression of UDP-Galactose Transporter Messenger RNA in Human Colon Cancer Tissues and Its Implication in Synthesis of Thomsen-Friedenreich Antigen and Sialyl Lewis A/X Determinants1. Cancer Res. 2001, 61, 4620–4627. [Google Scholar] [PubMed]
  212. Croci, D.O.; Cerliani, J.P.; Pinto, N.A.; GMorosi, L.; Rabinovich, G.A. Regulatory role of glycans in the control of hypoxia-driven angiogenesis and sensitivity to anti-angiogenic treatment. Glycobiology 2014, 24, 1283–1290. [Google Scholar] [CrossRef] [PubMed]
  213. Freitas, D.; Campos, D.; Gomes, J.; Pinto, F.; Macedo, J.A.; Matos, R.; Mereiter, S.; Pinto, M.T.; Polónia, A.; Gartner, F.; et al. O-glycans truncation modulates gastric cancer cell signaling and transcription leading to a more aggressive phenotype. EBioMedicine 2019, 40, 349–362. [Google Scholar] [CrossRef] [PubMed]
  214. Häuselmann, I.; Borsig, L. Altered tumor-cell glycosylation promotes metastasis. Front. Oncol. 2014, 4, 28. [Google Scholar] [CrossRef]
  215. Lau, K.S.; Partridge, E.A.; Grigorian, A.; Silvescu, C.I.; Reinhold, V.N.; Demetriou, M.; Dennis, J.W. Complex N-Glycan Number and Degree of Branching Cooperate to Regulate Cell Proliferation and Differentiation. Cell 2007, 129, 123–134. [Google Scholar] [CrossRef]
  216. Tominaga, N.; Hagiwara, K.; Kosaka, N.; Honma, K.; Nakagama, H.; Ochiya, T. RPN2-mediated glycosylation of tetraspanin CD63 regulates breast cancer cell malignancy. Mol. Cancer 2014, 13, 134. [Google Scholar] [CrossRef]
  217. Macedo-da-Silva, J.; Santiago, V.F.; Rosa-Fernandes, L.; Marinho, C.R.F.; Palmisano, G. Protein glycosylation in extracellular vesicles: Structural characterization and biological functions. Mol. Immunol. 2021, 135, 226–246. [Google Scholar] [CrossRef]
  218. Battke, C.; Ruiss, R.; Welsch, U.; Wimberger, P.; Lang, S.; Jochum, S.; Zeidler, R. Tumour exosomes inhibit binding of tumour-reactive antibodies to tumour cells and reduce ADCC. Cancer Immunol. Immunother. 2011, 60, 639–648. [Google Scholar] [CrossRef]
  219. Pigati, L.; Yaddanapudi, S.C.S.; Iyengar, R.; Kim, D.J.; Hearn, S.A.; Danforth, D.; Hastings, M.L.; Duelli, D.M. Selective release of microRNA species from normal and malignant mammary epithelial cells. PLoS ONE 2010, 5, e13515. [Google Scholar] [CrossRef]
  220. Smyth, T.; Petrova, K.; Payton, N.M.; Persaud, I.; Redzic, J.S.; Graner, M.W.; Smith-Jones, P.; Anchordoquy, T.J. Surface functionalization of exosomes using click chemistry. Bioconjug. Chem. 2014, 25, 1777–1784. [Google Scholar] [CrossRef]
  221. Gangoda, L.; Liem, M.; Ang, C.S.; Keerthikumar, S.; Adda, C.G.; Parker, B.S.; Mathivanan, S. Proteomic Profiling of Exosomes Secreted by Breast Cancer Cells with Varying Metastatic Potential. Proteomics 2017, 17, 1600370. [Google Scholar] [CrossRef] [PubMed]
  222. Menck, K.; Scharf, C.; Bleckmann, A.; Dyck, L.; Rost, U.; Wenzel, D.; Dhople, V.M.; Siam, L.; Pukrop, T.; Binder, C.; et al. Tumor-derived microvesicles mediate human breast cancer invasion through differentially glycosylated EMMPRIN. J. Mol. Cell Biol. 2015, 7, 143–153. [Google Scholar] [CrossRef] [PubMed]
  223. Koga, K.; Matsumoto, K.; Akiyoshi, T.; Kubo, M.; Yamanaka, N.; Tasaki, A.; Nakashima, H.; Nakamura, M.; Kuroki, S.; Tanaka, M.; et al. Purification, Characterization and Biological Significance of Tumor-derived Exosomes. Anticancer Res. 2005, 25, 3703–3707. [Google Scholar] [PubMed]
  224. Wolfers, J.; Lozier, A.; Raposo, G.; Regnault, A.; Théry, C.; Masurier, C.; Flament, C.; Pouzieux, S.; Faure, F.; Tursz, T.; et al. Tumor-derived exosomes are a source of shared tumor rejection antigens for CTL cross-priming. Nat. Med. 2001, 7, 297–303. [Google Scholar] [CrossRef] [PubMed]
  225. Nishida-Aoki, N.; Tominaga, N.; Kosaka, N.; Ochiya, T. Altered biodistribution of deglycosylated extracellular vesicles through enhanced cellular uptake. J. Extracell. Vesicles 2020, 9, 1713527. [Google Scholar] [CrossRef]
  226. Kavanagh, E.L.; Halasz, M.; Dowling, P.; Withers, J.; Lindsay, S.; Higgins, M.J.; Irwin, J.A.; Rudd, P.M.; Saldova, R.; McCann, A. N -Linked glycosylation profiles of therapeutic induced senescent (TIS) triple negative breast cancer cells (TNBC) and their extracellular vesicle (EV) progeny. Mol. Omics 2021, 17, 72–85. [Google Scholar] [CrossRef]
  227. Lima, L.G.; Ham, S.; Shin, H.; Chai, E.P.Z.; Lek, E.S.H.; Lobb, R.J.; Müller, A.F.; Mathivanan, S.; Yeo, B.; Choi, Y.; et al. Tumor microenvironmental cytokines bound to cancer exosomes determine uptake by cytokine receptor-expressing cells and biodistribution. Nat. Commun. 2021, 12, 3543. [Google Scholar] [CrossRef]
  228. Tan, Z.; Cao, L.; Wu, Y.; Wang, B.; Song, Z.; Yang, J.; Cheng, L.; Yang, X.; Zhou, X.; Dai, Z. Bisecting GlcNAc modification diminishes the pro-metastatic functions of small extracellular vesicles from breast cancer cells. J. Extracell. Vesicles 2020, 10, e12005. [Google Scholar] [CrossRef]
  229. Hait, N.C.; Maiti, A.; Wu, R.; Andersen, V.L.; Hsu, C.-C.; Wu, Y.; Chapla, D.G.; Takabe, K.; Rusiniak, M.E.; Bshara, W.; et al. Extracellular sialyltransferase st6gal1 in breast tumor cell growth and invasiveness. Cancer Gene Ther. 2022, 29, 1662–1675. [Google Scholar] [CrossRef]
  230. Stowell, S.R.; Ju, T.; Cummings, R.D. Protein glycosylation in cancer. Annu. Rev. Pathol. 2015, 10, 473–510. [Google Scholar] [CrossRef]
  231. Toraskar, J.; Magnussen, S.N.; Hagen, L.; Sharma, A.; Hoang, L.; Bjørkøy, G.; Svineng, G.; Steigedal, T.S. A novel truncated form of nephronectin is present in small extracellular vesicles isolated from 66cl4 cells. J. Proteome Res. 2019, 18, 1237–1247. [Google Scholar] [CrossRef]
  232. Cao, L.; Wu, Y.; Wang, X.; Li, X.; Tan, Z.; Guan, F. Role of Site-Specific Glycosylation in the I-Like Domain of Integrin β1 in Small Extracellular Vesicle-Mediated Malignant Behavior and FAK Activation. Int. J. Mol. Sci. 2021, 22, 1770. [Google Scholar] [CrossRef]
  233. AL-Abedi, R.; Tuncay Cagatay, S.; Mayah, A.; Brooks, S.A.; Kadhim, M. Ionising Radiation Promotes Invasive Potential of Breast Cancer Cells: The Role of Exosomes in the Process. Int. J. Mol. Sci. 2021, 22, 11570. [Google Scholar] [CrossRef]
  234. Bard, M.P.; Hegmans, J.P.; Hemmes, A.; Luider, T.M.; Willemsen, R.; Severijnen, L.A.A.; Van Meerbeeck, J.P.; Burgers, S.A.; Hoogsteden, H.C.; Lambrecht, B.N. Proteomic analysis of exosomes isolated from human malignant pleural effusions. Am. J. Respir. Cell Mol. Biol. 2004, 31, 114–121. [Google Scholar] [CrossRef] [PubMed]
  235. 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. 2018, 21, 190–202. [Google Scholar] [CrossRef] [PubMed]
  236. Vergauwen, G.; Tulkens, J.; Pinheiro, C.; Avila Cobos, F.; Dedeyne, S.; De Scheerder, M.A.; Vandekerckhove, L.; Impens, F.; Miinalainen, I.; Braems, G.; et al. Robust sequential biophysical fractionation of blood plasma to study variations in the biomolecular landscape of systemically circulating extracellular vesicles across clinical conditions. J. Extracell. Vesicles 2021, 10, e12122. [Google Scholar] [CrossRef]
  237. Gonzalez, E.; Piva, M.; Rodriguez-Suarez, E.; Gil, D.; Royo, F.; Elortza, F.; Falcon-Perez, J.M.; Vivanco, M.D.M. Human Mammospheres Secrete Hormone-Regulated Active Extracellular Vesicles. PLoS ONE 2014, 9, e83955. [Google Scholar] [CrossRef]
  238. 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]
  239. 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]
  240. Moon, P.G.; Lee, J.E.; Cho, Y.E.; Lee, S.J.; Jung, J.H.; Chae, Y.S.; Bae, H.I.; Kim, Y.B.; Kim, I.S.; Park, H.Y.; et al. Identification of Developmental Endothelial Locus-1 on Circulating Extracellular Vesicles as a Novel Biomarker for Early Breast Cancer Detection. Clin. Cancer Res. 2016, 22, 1757–1766. [Google Scholar] [CrossRef]
  241. Xu, G.; Huang, W.; Du, S.; Huang, M.; Lyu, J.; Zhou, F.; Zhu, R.; Cao, Y.; Xv, J.; Li, N.; et al. Proteomic Characterization of Serum Small Extracellular Vesicles in Human Breast Cancer. bioRxiv 2021, bioRxiv:11.26.470104. [Google Scholar] [CrossRef]
  242. SUN, J.G.; LI, X.B.; YIN, R.H.; LI, X.F. lncRNA VIM-AS1 promotes cell proliferation, metastasis and epithelial–mesenchymal transition by activating the Wnt/β-catenin pathway in gastric cancer. Mol. Med. Rep. 2020, 22, 4567–4578. [Google Scholar] [CrossRef] [PubMed]
  243. Wu, S.; Du, Y.; Beckford, J.; Alachkar, H. Upregulation of the EMT marker vimentin is associated with poor clinical outcome in acute myeloid leukemia. J. Transl. Med. 2018, 16, 170. [Google Scholar] [CrossRef] [PubMed]
  244. Ozawa, P.M.M.; Vieira, E.; Lemos, D.S.; Souza, I.L.M.; Zanata, S.M.; Pankievicz, V.C.; Tuleski, T.R.; Souza, E.M.; Wowk, P.F.; Urban, C.D.A.; et al. Identification of miRNAs Enriched in Extracellular Vesicles Derived from Serum Samples of Breast Cancer Patients. Biomolecules 2020, 10, 150. [Google Scholar] [CrossRef] [PubMed]
  245. 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]
  246. Cai, G.X.; Lin, L.; Zhai, X.M.; Guo, Z.W.; Wu, Y.S.; Ye, G.L.; Liu, Q.; Chen, L.S.; Xing, G.Y.; Zhao, Q.H.; et al. A plasma-derived extracellular vesicle mRNA classifier for the detection of breast cancer. Gland Surg. 2021, 10, 2002009. [Google Scholar] [CrossRef]
  247. 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, 1823–1834. [Google Scholar] [CrossRef]
  248. Su, Y.; Li, Y.; Guo, R.; Zhao, J.; Chi, W.; Lai, H.; Wang, J.; Wang, Z.; Li, L.; Sang, Y.; et al. Plasma extracellular vesicle long RNA profiles in the diagnosis and prediction of treatment response for breast cancer. NPJ Breast Cancer 2021, 7, 154. [Google Scholar] [CrossRef]
  249. Tang, S.; Zheng, K.; Tang, Y.; Li, Z.; Zou, T.; Liu, D. Overexpression of serum exosomal HOTAIR is correlated with poor survival and poor response to chemotherapy in breast cancer patients. J. Biosci. 2019, 44, 37. [Google Scholar] [CrossRef]
  250. Zhong, G.; Wang, K.; Li, J.; Xiao, S.; Wei, W.; Liu, J. Determination of Serum Exosomal H19 as a Noninvasive Biomarker for Breast Cancer Diagnosis. Onco. Targets Ther. 2020, 13, 2563–2571. [Google Scholar] [CrossRef]
  251. Wu, M.; Wang, M.; Jia, H.; Wu, P. Extracellular vesicles: Emerging anti-cancer drugs and advanced functionalization platforms for cancer therapy. Drug Deliv. 2022, 29, 2513–2538. [Google Scholar] [CrossRef] [PubMed]
  252. Liu, C.; Wang, Y.; Li, L.; He, D.; Chi, J.; Li, Q.; Wu, Y.; Zhao, Y.; Zhang, S.; Wang, L.; et al. Engineered extracellular vesicles and their mimetics for cancer immunotherapy. J. Control. Release 2022, 349, 679–698. [Google Scholar] [CrossRef] [PubMed]
  253. Ahmadi, M.; Hassanpour, M.; Rezaie, J. Engineered extracellular vesicles: A novel platform for cancer combination therapy and cancer immunotherapy. Life Sci. 2022, 308, 120935. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Challenges and perspectives regarding omics research on BC-derived EVs. This review presents the current state of the art of the most-studied omics topics on EVs from BC, namely, transcriptomics and proteomics. However, we must fill omics-related gaps before proposing reliable EV-based tools for this disease. Here, we cited some challenges for future research. As vesicles are heterogeneous in terms of size, biogenesis, and cargo, authors must standardize the reporting of methods for the isolation, quantification, characterization, and profiling of EVs. Furthermore, consistent findings in relation to EVs are characterized by their ability to be replicated. Nevertheless, many studies use targeted analysis approaches, which can bias observations. In addition, such replicability must be related to characterizing different individuals of the same subgroup or cell lines of the same subtype. To evaluate this correctly, it is necessary to increase the number of studies comparing less-studied BC cell lines and include a translational approach between tumor cell markers and their vesicular pairs. Regarding associations with the subtype of BC patients, there are gaps produced by the lack of available information about the molecular or clinical profiles of these patients, which can complicate future secondary analysis. After conquering this challenge, we can promisingly combine data from different omics studies of BC-derived EVs and select potentially tumor-derived EVs via liquid biopsies from patients to debug or edit these vesicles and induce a beneficial effect in BC patients. Image created on BioRender.com.
Figure 1. Challenges and perspectives regarding omics research on BC-derived EVs. This review presents the current state of the art of the most-studied omics topics on EVs from BC, namely, transcriptomics and proteomics. However, we must fill omics-related gaps before proposing reliable EV-based tools for this disease. Here, we cited some challenges for future research. As vesicles are heterogeneous in terms of size, biogenesis, and cargo, authors must standardize the reporting of methods for the isolation, quantification, characterization, and profiling of EVs. Furthermore, consistent findings in relation to EVs are characterized by their ability to be replicated. Nevertheless, many studies use targeted analysis approaches, which can bias observations. In addition, such replicability must be related to characterizing different individuals of the same subgroup or cell lines of the same subtype. To evaluate this correctly, it is necessary to increase the number of studies comparing less-studied BC cell lines and include a translational approach between tumor cell markers and their vesicular pairs. Regarding associations with the subtype of BC patients, there are gaps produced by the lack of available information about the molecular or clinical profiles of these patients, which can complicate future secondary analysis. After conquering this challenge, we can promisingly combine data from different omics studies of BC-derived EVs and select potentially tumor-derived EVs via liquid biopsies from patients to debug or edit these vesicles and induce a beneficial effect in BC patients. Image created on BioRender.com.
Ijms 24 13022 g001
Figure 2. Extracellular vesicle (EV) miRNAs in breast-cancer-related studies. Sankey plots show the number of studies mentioning each relevant vesicular miRNA from cell supernatant (A) or human bodily fluids (B). The cell lines in which the EV cargo was analyzed are classified into the main BC subtypes following the criteria given in Dai et al.’s (2017) study [124]. For studies on EVs collected from BC patients, the subtype information was retrieved from each study. H: Her2, TNA: Triple-Negative A, TNB: Triple-Negative B, LA: Luminal A, and LB: Luminal B.
Figure 2. Extracellular vesicle (EV) miRNAs in breast-cancer-related studies. Sankey plots show the number of studies mentioning each relevant vesicular miRNA from cell supernatant (A) or human bodily fluids (B). The cell lines in which the EV cargo was analyzed are classified into the main BC subtypes following the criteria given in Dai et al.’s (2017) study [124]. For studies on EVs collected from BC patients, the subtype information was retrieved from each study. H: Her2, TNA: Triple-Negative A, TNB: Triple-Negative B, LA: Luminal A, and LB: Luminal B.
Ijms 24 13022 g002
Figure 3. Relevant putative markers in BC-derived EVs. Breast cancer cells produce a great diversity of EVs. These EVs can be classified into subpopulations based on their proteomic and transcriptomic cargo. In this review, we associate some BC-derived EV subpopulations with tumor-related functions. In addition, we include putative markers related to their types (miRNA, lncRNA, mRNA, circRNA, or protein) for each subpopulation. Image created on BioRender.com.
Figure 3. Relevant putative markers in BC-derived EVs. Breast cancer cells produce a great diversity of EVs. These EVs can be classified into subpopulations based on their proteomic and transcriptomic cargo. In this review, we associate some BC-derived EV subpopulations with tumor-related functions. In addition, we include putative markers related to their types (miRNA, lncRNA, mRNA, circRNA, or protein) for each subpopulation. Image created on BioRender.com.
Ijms 24 13022 g003
Figure 4. Extracellular vesicle (EVs) proteins in breast cancer proteome studies. (A) The most frequently identified proteins in the evaluated studies. The donut graph shows the corresponding subcellular locations. (B) Subcellular localization of proteins identified in EVs in at least two studies. (C) Cell-cycle-related and p53 pathways in which EV proteins participate.
Figure 4. Extracellular vesicle (EVs) proteins in breast cancer proteome studies. (A) The most frequently identified proteins in the evaluated studies. The donut graph shows the corresponding subcellular locations. (B) Subcellular localization of proteins identified in EVs in at least two studies. (C) Cell-cycle-related and p53 pathways in which EV proteins participate.
Ijms 24 13022 g004
Table 1. Protein clusters found in EVs derived from cell lines and bodily fluids from humans with breast cancer.
Table 1. Protein clusters found in EVs derived from cell lines and bodily fluids from humans with breast cancer.
Protein ClustersHighlightBC-Related Subtype 1 (Cell Lines)
or Type of Sample (Bodily Fluids)
Reference
PDIA1, PDIA3, SUSD2,
LGALS3BP, PRKCSH
The SKBR3 cell line presents a higher number of peptide spectral matches with invasion-related proteins compared to MCF7 and MCF10.Her2 and Luminal A[186]
FZD6, DVL1, PK1, VANGL1Inhibition of Fzd6, Dvl1, Pk1, and Vangl1 (PCP pathway) reduced tumor cell motility.Triple-Negative B[238]
VIM, LGALS3BP, ANXA1,
EDIL3, FLNB, TGM2, CTSD,
PLAU, PRSS23, SERPINE1,
CLIC1, EPHA2, NDRG1,
HSPA9, HSP90AB1, SRP68
There are relatively few proteins that
are differentially expressed in exosomes from MDA-MB-231 cells (invasive/metastatic) and MCF-7 cells (non-invasive/non-metastatic). However, the identified proteins that were upregulated at MDA-MB-231 have known functions in migration.
Luminal A and Triple-Negative B[14]
C4B, EDIL3, VTN, ANXA6,
NUMA1, FLNC, TXNRD1,
COL5A1, PSMB4, FLNA,
EEA1, ENPP1, PSMA1, ILF2,
RPL12, ISG15, PSMB7,
PON1, PSMD6, LAMP1, FTL,
PSMC5, RPL27
EVs from young women breast cancer patients drive
increased invasion of non-malignant cells via the Focal
Adhesion Kinase pathway. These results suggest that the protein content of EVs from MDA-MB231 and MCF10DCIS.com cells reflects the biologic differences between these invasive and non-invasive breast cancer cells.
Triple-Negative B[171]
ZEB1, SNAI1EVs from metastatic cells can affect the behavior of less-aggressive neighboring cells.Luminal A and Triple-Negative B[237]
ADAM10Exosomal ADAM10 increases aldehyde dehydrogenase expression in breast cancer cells through Notch receptor activation and enhances motility through the GTPase RhoA.Triple-Negative B[239]
ADAM10ADAM-10 expression levels are increased in exosomes from luminal cancer subtype blood in comparison with healthy patients.Blood[173]
PEPD, NCL, PARP1, ACTA2,
ACTG2, TBCA, TTYH3,
MATR3, KPNB1, KRT16,
RANBP2, CCT6A
The application of protein signatures to discriminate breast cancer patients with or without metastasis yielded a sensitivity of 81% and a specificity of 81%.Serum[241]
CD151CD151 levels in serum exosomes derived from TNBC were significantly higher than the levels of exosomes from healthy individuals. This protein facilitates the secretion of ribosomal proteins via exosomes while inhibiting the secretion of complement proteins. CD151-deleted exosomes significantly decreased cell migration and invasion.Serum[175]
ADAM9, AXL, TIMP2, PLAU,
VIM, ITGA5, EPHA2, NRP1,
PABPC1, DNAJA1
TNBC-specific signature proteins featured prominently in angiogenesis, cell motility and cell migration, and integrin binding.All main subtypes[176]
A1BG, ACSM3, ADAM10,
AHSG, AMBP, APOA1,
APPBP2, BANF1, BMP7,
C3, CLK3, ELL2, GTSE1,
HIG1AN, HMOX1, IGF2R,
ITIH4, KRT1, KRT6A, KRT6B,
MBD4, PCNT, PIBF1, PPM1A,
SERPINA1, SERPINB7
Exosomal proteins from the blood of breast cancer patients are associated with crucial steps of tumor progression and metastasis.Blood[177]
ANXA5, ANXA6, VIM, CD44,
EIF4A1, ITGA2, CD147, ENO1,
ITGB1, FLNB, FLNA, EGFR
It was observed that the metastatic proteins ENO1 and ITGB3 were upregulated in all biological replicates of MDA-MB-231 EVs.Luminal A and
Triple-Negative B
[179]
GLUT-1, ADAM10, GPC-1GPC-1, ADAM10, and GLUT-1 proteins may be novel potential biomarkers for breast cancer detection and prognosis.Luminal A and
Triple-Negative B
[181]
1 Cell lines were classified into subtypes using the data reported in study conducted by Dai et al. (2017) [124].
Table 2. Features of vesicular cargo conducive to their use as potential biomarkers for breast cancer.
Table 2. Features of vesicular cargo conducive to their use as potential biomarkers for breast cancer.
MarkersRNA TypeMarker TypeComparison
(Group 1 vs. Group 2)
SourceMethod for
EV Isolation
Total or Specific
Group of EVs
Method for
Vesicular RNA
Isolation
Method for
Vesicular RNA Quantification
Area under
the Curve
Ref.
hsa-let-7a-5p and hsa-miR-222-3pmiRNAClassificationInflammatory BC (n = 23) vs Non-Inflammatory BC (n = 34)PlasmaCombination of methodsTotalOrganic extraction methodqRT-PCR0.973[100]
hsa-miR-142-5pmiRNAClassificationTNBC (n = 15) vs. Luminal A (n = 16)SerumPrecipitation using chemical reagentTotalOrganic extraction methodqRT-PCR0.921[244]
hsa-miR-4448, hsa-miR-2392, hsa-miR-2467-3p and
hsa-miR-4800-3p
miRNAClassificationrecurrent TNBC (n = 12) vs. Non recurrent TNBC (n = 12)SerumPrecipitation using chemical reagentTotalOrganic extraction methodqRT-PCR0.765[245]
HLA-DRB1, HAVCR1, ENPEP, TIMP1, CD36, MARCKS, DAB2, and CXCL14mRNADiagnosisBC patients (n = 101) vs.
Healthy individuals or those with benign breast disease (n = 81)
PlasmaAffinity-based binding to spin columnsTotalOrganic extraction methodqRT-PCR0.77[246]
PGR, ERBB2mRNADiagnosisBC patients (n = 63) vs.
Healthy individuals (n = 20)
PlasmaUltracentrifugationTotalAffinity-based binding to spin columnsddPCR0.93[35]
hsa-miR-424, hsa-miR-423, hsa-miR-660, and
hsa-let-7i
miRNADiagnosisBC patients (n = 69) vs.
Healthy individuals (n = 40)
UrineSize-exclusion filterTotalSilicon-carbide-based methodqRT-PCR0.995[119]
hsa-miR-21miRNADiagnosisBC patients (n = 30) vs.
Healthy individuals or other tumors (n = 54)
PlasmaUltracentrifugationTotalOrganic extraction methodqRT-PCR0.961[39]
hsa-miR-16, hsa-miR-21, hsa-miR-429, and hsa-miR-9miRNADiagnosisBC patients (n = 62) vs.
Healthy individuals (n = 20)
PlasmaMicrofluidic deviceCD49f and EpCAM-positive EVsOrganic extraction methodqRT-PCR0.880[106]
hsa-miR-142-5p, hsa-miR-320a, and hsa-miR-4433b-5pmiRNADiagnosisBC patients (n = 31) vs.
Healthy individuals (n = 16)
SerumPrecipitation using chemical reagentTotalOrganic extraction methodqRT-PCR0.839[244]
hsa-miR-421miRNADiagnosisBC patients (n = 20) vs.
Healthy individuals (n = 10)
PlasmaUltracentrifugationTotalOrganic extraction methodqRT-PCR0.835[115]
hsa-miR-421miRNADiagnosisstage I BC patients (n = 13) vs.
Healthy individuals (n = 10)
PlasmaUltracentrifugationTotalOrganic extraction methodqRT-PCR0.809[115]
hsa-miR-375, hsa-miR-655-3p, hsa-miR-548b-5p, and hsa-miR-24-2-5pmiRNADiagnosisstage I BC patients (n = 12) vs.
Healthy individuals (n = 10)
PlasmaPrecipitation using chemical reagentTotalOrganic extraction methodRNA-seq0.808[104]
hsa-miR-17-5pmiRNADiagnosisBC patients (n = 83) vs.
Healthy individuals (n = 34)
SerumNANANAqRT-PCR0.784[247]
hsa-miR-1246 and hsa-miR-21miRNADiagnosisBC patients (n = 16) vs.
Healthy individuals (n = 16)
PlasmaCombination of methodsTotalOrganic extraction methodRNA-seq0.727[38]
BEX2, AC104843.1, AL136981.2, KRT19, NPM1P25, CTSG, CBR3, HOXB7, AL691447.3, RNA5SP141, and circRNA chr13_42953948_42970670_-lncRNA /circRNADiagnosisstages I-II BC patients (n = 63) vs.
Healthy individuals or those with benign breast disease (n = 60)
PlasmaAffinity-based binding to spin columnsTotalOrganic extraction methodRNA-seq0.940[248]
HOTAIRlncRNADiagnosisBC patients (n = 15) vs.
Healthy individuals (n = 15)
SerumCombination of methodsTotalAffinity-based binding to spin columnsqRT-PCR0.918[249]
H19lncRNADiagnosisBC patients (n = 50) vs.
Healthy individuals or those with benign breast disease (n = 100)
SerumPrecipitation using chemical reagentTotalOrganic extraction methodqRT-PCR0.870[250]
hsa-miR-150-5pmiRNAPrognosisRecurrent BC (n = 12) vs. non-recurrent BC (n = 15)PlasmaPrecipitation using chemical reagentTotalOrganic extraction methodRNA-seq0.705[132]
hsa-miR-1246 and hsa-miR-155miRNATreatment Response
Prediction
Tratuzumab-resistant (n = 32) vs. Trastuzumab-sensitive BC patients (n = 36)PlasmaPrecipitation using chemical reagentTotalMagnetic-beads-based methodqRT-PCR0.898[98]
MSMO1lncRNATreatment Response
Prediction
Patients with PCR (n = 24) vs patients with no-PCR after treatment with NACT (n = 34)PlasmaAffinity-based binding to spin columnsTotalOrganic extraction methodRNA-seq0.79[248]
Ref.: Reference; BC: Breast Cancer; TNBC: Triple-Negative Breast Cancer; EV: Extracellular Vesicle; PCR: pathological complete response; NACT: neoadjuvant chemotherapy; NA: Not Available.
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

Murillo Carrasco, A.G.; Otake, A.H.; Macedo-da-Silva, J.; Feijoli Santiago, V.; Palmisano, G.; Andrade, L.N.d.S.; Chammas, R. Deciphering the Functional Status of Breast Cancers through the Analysis of Their Extracellular Vesicles. Int. J. Mol. Sci. 2023, 24, 13022. https://doi.org/10.3390/ijms241613022

AMA Style

Murillo Carrasco AG, Otake AH, Macedo-da-Silva J, Feijoli Santiago V, Palmisano G, Andrade LNdS, Chammas R. Deciphering the Functional Status of Breast Cancers through the Analysis of Their Extracellular Vesicles. International Journal of Molecular Sciences. 2023; 24(16):13022. https://doi.org/10.3390/ijms241613022

Chicago/Turabian Style

Murillo Carrasco, Alexis Germán, Andreia Hanada Otake, Janaina Macedo-da-Silva, Veronica Feijoli Santiago, Giuseppe Palmisano, Luciana Nogueira de Sousa Andrade, and Roger Chammas. 2023. "Deciphering the Functional Status of Breast Cancers through the Analysis of Their Extracellular Vesicles" International Journal of Molecular Sciences 24, no. 16: 13022. https://doi.org/10.3390/ijms241613022

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