Skip to main content

The emerging roles and therapeutic potential of exosomes in epithelial ovarian cancer

Abstract

Ovarian cancer (OC) is one of the three types of malignant tumors in the female reproductive system, and epithelial ovarian cancer (EOC) is its most typical form. Due to the asymptomatic nature of the early stages and resistance to chemotherapy, EOC has both a poor prognosis and a high fatality rate. Current treatments for OC are very limited, and the 5-years survival rate is approximately 30%. Exosomes, which are microvesicles ranging from approximately 30–100 nm in size that are secreted by living cells, can be produced from different cell types and detected in various body fluids. Cancer cells can secrete more exosomes than healthy cells, and more importantly, the content of cancer cell-derived exosomes is distinct. The exosomes shedding from tumor cells are considered to be involved in tumor progression and metastasis. As such, exosomes are expected to be potential tools for tumor diagnosis and treatment. In this review, we briefly present the emerging roles of exosomes in OC and summarize related articles about their roles as diagnostic or prognostic biomarkers and in the treatment and drug resistance of OC.

Background

Epithelial ovarian cancer (EOC) is one of the most malignant tumors in the female reproductive system. As cancer statistics in China [1] revealed, the mortality rate of ovarian cancer (OC) has been rising for the past few years (comparing 2001–2003 with 2003–2011). In addition, this trend also existed in the United States, and the estimated mortality rates of OC would rank as the fifth highest based on existing data (2007–2011) [2]. Unlike other tumors, OC tumors tend to metastasize within the peritoneum but rarely disseminate through the vasculature [3], suggesting that OC may require treatment that is different from that for other tumors. Current clinical treatments for OC are very limited, and the surgery (cytoreduction or debulking), chemotherapy or treatment with a novel agent are the main therapeutic options. More than 75% of women have advanced disease (stage III or IV) at the time of diagnosis, that the 5-years survival rate among these patients is less than 25% [4]. Due to the asymptomatic nature of this disease during the early stages and resistance to chemotherapy, a poor prognosis can be expected [5].

Exosomes were proposed as products of membrane exfoliation at first. Johnstone first described and named them “exosomes” in 1987, and he suggested that exosomes might have specific membrane-related functions [6]. The potential primary function of exosomes is intercellular communication. Exosomes may have distinct biological activities, and their components depend on the cellular origin and are highly variable [7]. Exosomes contain many proteins, mRNAs, miRNAs, DNAs, lipids and transcriptional factors. In addition, a multitude of pathways can be activated because of cellular interactions with exosomal molecules, including mRNAs, miRNAs and proteins (e.g., heat shock proteins [HSPs] and adhesion molecules) [8].

It has been gradually accepted that cancer cells secrete more exosomes than healthy cells. Moreover, the contents of these exosomes have been found distinct from each other [9]. Remarkably, cancer cell-derived exosomes can provide a suitable microenvironment for cancer development, such as cell proliferation [10,11,12], drug resistance [13], angiogenesis and metastasis [14], immune modulation [15] and premetastatic niche formation [16]. In addition, these features are all present in OC. Therefore, exosomes are expected to be potential diagnostic, prognostic, and even therapeutic tools for EOC. Certainly, such applications are still under development and far from clinical translation. Exciting research has shown that exosomes can serve as drug delivery systems, similar to nanoparticles. Scientists loaded exosomes with chemotherapeutic drugs, which can be used to kill tumor cells effectively without typical side effects [17]. Based on current nanoparticle research, such drug delivery systems will have more applicability in clinic.

Although the field of exosomes is relatively new, they are involved in OC and expected to have applications in OC as well. In this review, we briefly present the emerging roles of exosomes in OC and summarize related articles about their roles as diagnostic or prognostic biomarkers and in the treatment and drug resistance of OC, especially EOC.

Exosomes as biomarkers of OC

The biomarker mainly used for the diagnosis and prognosis of EOC is CA125, which has low sensitivity and specificity. Recent studies have revealed that exosomes can transport enzymes from EOC cells. Furthermore, an increased level of exosomes in the peritoneal fluid has been found to be correlated with tumor progression [18, 19], indicating that the level of exosomes in the blood is correlated with the stage of OC. Thus, exosomes themselves and their contents have potential as tumor-specific markers. Although no exosomes-related indicator comparable with CA125 has been widely accepted. In addition, the lack of large-scale clinical trials and the immature purification and detection methods remain unresolved issues. The writers still believe that exosomes are promising as biomarkers of OC.

Exosomal miRNAs as diagnostic and prognostic biomarkers

Recent studies have shown that the role of exosomes was protecting miRNAs from RNases, which created conditions for using exosomal miRNAs to diagnose OC [20]. MiRNA levels are important and often detected. Eight particular miRNAs (e.g., miR-21) in exosomes isolated from the serum of women with benign disease or various stages of EOC demonstrated diagnostic promise, because they were not detectable in the normal controls [21]. Those findings indicate that miRNAs of circulating tumor-derived exosomes could potentially be used as surrogate diagnostic markers for biopsy profiling or extended to screening asymptomatic populations. A similar phenomenon was observed in miR-222-3p, compared with healthy individuals, the expression of exosomal miR-222-3p was significantly increased in the serum of EOC patients [22]. MiR-21 was involved in the oncogenesis of ovarian serous carcinoma (OSC). The exosomal transfer of miR-21 could promote oncogenic transformation in target cells distant from the primary tumor, and it could be used as a diagnostic tool [23]. The high concentrations of exosomes in EOC patients suggested excessive and active exosomal secretion. In a cohort of 163 EOC patients, the levels of miR-200b and miR-200c were higher in International Federation of Gynecologists and Obstetricians (FIGO) stage III-IV patients, including those with lymph node metastasis, than in FIGO stage I-II patients [24]. This analysis showed that these miRNAs were highly suitable for tumor staging, diagnosis and prognosis.

Exosomes in the urine have gained attention, and urinary exosomal miRNAs are easily available samples that have been recently explored, particularly in urological and gynecological diseases. Zavesky confirmed a significant up-regulation of miR-92a in OC urinary samples, which could use for diagnostic purposes [25]. In addition, miRNA microarray data showed that miR-30a-5p was up-regulated in the urine samples of ovarian serous adenocarcinoma (OSA) patients compared with healthy controls. Furthermore, lower urine levels of miR-30a-5p were found in gastric cancer (GC) and colon carcinoma (CC) patients, suggesting that urinary miR-30a-5p may be particular to OC and demonstrating the use of exosomal urinary miR-30a-5p as a specific diagnostic marker [26]. However, the diagnostic potential of exosomal urinary miRNAs may depend on the urine fraction and might be vulnerable to the external environment, which is an obvious constraint.

Considering that the sources of exosomal miRNAs were very extensive, and differed from one another, the type and expression level of miRNAs was very different. Moreover, the various studies mentioned before had gathered information on patients from different type of ovarian cancer, thus they presented different characteristics in the number of patients, varied specimen and quantity. Despite all the research with one table list, it was lack of inconsistency between different studies.

Exosome-related proteins as biomarkers

In this section, we summarize exosome-related proteins, such as endosomal membrane proteins, HSPs, enzymes, and globulins. These proteins are involved in adhesion, angiogenesis, proliferation, and metabolism [27]. In-depth proteomic analyses of OC cell line-derived exosomes revealed the differential enrichment of functional proteome categories [28]. When examining the protein profile of purified exosomes derived from two OC cell lines (OVCAR-3 and IGROV1), 1017 proteins were identified in both exosome types, including all the major exosomal protein markers. These data indicated that exosomes might serve as a resource of blood-based protein biomarkers in OC [29]. Comparing malignant and cirrhotic ascites, a proteomic analysis identified 424 proteins specific to malignant ascites. In addition, a functional analysis demonstrated that the major differences between them were the clusters of spliceosome proteins. The OC ascites proteome may provide markers for predicting patient survival and the effectiveness of different chemotherapy methods [30]. OC patient’s plasma contains higher levels of exosomal proteins than that of benign tumor patients or healthy controls. Significantly, high exosomal protein levels were observed in both newly diagnosed OC patients and those with advanced stages of OC. In addition, the levels varied when examined during and after chemotherapy [31]. As a result, changes in exosomal protein levels might be useful for predicting responses to therapy and prognosis in OC patients when combined with clinical data.

Epithelial cell adhesion molecule (EpCAM) overexpression correlated with an escalation in epithelial cell proliferation during tumor development [32]. The exosomal EpCAM levels in normal individuals were very low, while an increase in the exosomal EpCAM level was found to be related to the stage and severity of ovarian carcinoma. These findings indicated that EpCAM could be useful for the diagnosis of OC with the detection of tumor-derived exosomes [21]. CD24 is an established marker of poor prognosis in ovarian and other types of carcinomas, and it has been identified in exosomes isolated from the ascites fluid of ovarian carcinoma patients. What is more, most of the CD24-positive exosomes are secreted by tumor cells [33]. EpCAM and CD24 in ovarian tumor-derived exosomes are promising alternatives for early detection [34]. In addition, they can distinguish OC-derived exosomes from vesicles originating from normal cells. Furthermore, as shown in a small set of pre- and post-treatment OC patient ascites samples, changes in the exosomal EpCAM and CD24 levels can accurately classify patients as being responsive or nonresponsive to therapy [35, 36], which has great value for evaluating prognosis. Activated leukocyte cell adhesion molecule (ALCAM) is involved in cell-cell interactions in cancer. Serum ALCAM levels have also been found to be significantly higher in EOC patients than in controls, and full-length transmembrane ALCAM was detected in exosomes extracted from EOC ascites samples. Studies have shown that ALCAM is a marker of EOC and correlates with more aggressive tumors [37].

Tumor-reactive immunoglobulins can also be used as diagnostic markers in OC. Serum samples from OC patients exhibited significantly greater immunoreactivity levels than samples from either normal controls or women with benign disease. Analyzing IgG recognition of specific exosomal antigens can distinguish benign ovarian masses from OC, as well as early- and late-stage OC. The quantitative assessment of IgG reactivity with tumor-derived exosomal proteins can be used as diagnostic markers for OC [38]. In addition, claudin proteins frequently overexpressed in OCs, and full-length claudins can be found in OC cell-derived exosomes. Moreover, nearly half of the plasma samples from OC patients exhibited the presence of claudin-4-containing exosomes [39]. Thus, claudin-4 can be used for the detection of OC.

A proteomics analysis of exosomes derived from two late-stage OC cell lines demonstrated that the pentose phosphate pathway was a dominant mechanism in exosome-mediated intracellular communication. Glucose-6-phosphate dehydrogenase (G6PD), transketolase (TK) and transaldolase 1 (TA1) are three key enzymes in the regulation of this pathway and may become diagnostic or prognostic biomarkers of OC [40]. The inhibition of N-glycosylation processing caused changes in the composition of extracellular vesicles and induced a decrease in the levels of several glycoproteins, which showed that the identified glycosignatures of extracellular vesicles could serve as novel biomarkers for OC [41].

Yong Zeng et al. found that the ExoSearch chip offers a method for the isolation and release of blood/plasma exosomes in various preparation volumes (10 μL to 10 mL). In addition, the blood-based diagnosis of OC via three exosomal protein markers for tumor (CA125, EpCAM, and CD24) showed significant diagnostic power [42] in that tumor testing with a drop of blood was possible. All biomarkers related to exosomes are summarized in Table 1

Table 1 Exosomes serve as biomarkers of OC

Exosomes as prospctive therapeutic targets in OC

Exosomes are related to the immune system, which may provide the foundation for future immunotherapies

The limitations of current chemotherapies for OC have motivated the development of new treatments. Evidence has shown that exosomes released by a tumor can suppress the immune system of a patient and prepare niches for metastatic spread [43]. Cancer-related exosomes exhibit powerful influences that benefit tumors via a variety of biological mechanisms. Understanding the underlying and complex exosome-mediated immunosuppressive mechanisms is important to prevent the immune escape of tumor cells and identify novel treatments for cancer. The ability to block immunosuppressive factors in tumor microenvironments is expected to enhance antitumor immune responses in OC patients.

Myeloid-derived suppressor cells (MDSCs), whose main feature is the potent immunosuppressive effect, are a major component of the tumor microenvironment. MDSCs are a heterogeneous group that includes macrophages, tumor-associated macrophages (TAMs), dendritic cells (DCs), monocytes, and granulocytes [44]. Dendritic and lymphoid exosomes can regulate immune activation. Tumors can release exosomes mimicking membranous material, resulting in the deletion of reactive lymphocytes. In addition, these exosomes expressed many reactive molecules, which suppressed the expression of T cell activation signal components, thus suppressing the immune activity of T cells and inducing apoptosis [45]. Exosomes isolated from the malignant ascites of OC patients could be internalized by monocytes and induced the production of cytokines. Exosomes from ascites can trigger toll-like receptor (TLR)-dependent signaling pathways in monocyte-precursor cells, which are important for the induction of immunosuppressive mechanisms during cancer progression [46]. The function and differentiation of macrophages can also be adjusted by exosomes. EOC-derived exosomes can activate macrophages to exhibit a TAM-like phenotype, which can facilitate the progression of cancer. MiR-222-3p can transfer from EOC cells to macrophages by exosomes, thereby effectively regulating the polarization of tumor-promoting M2 macrophages [22]. In addition, EOC exosomes have shown to be involved in the immunosuppression of natural killer (NK) cells, which are confirmed to be important in two cytotoxic pathways for anticancer immunity, the NK C-type lectin-like receptor (NKG2D) receptor-ligand pathway and the CD226/DNAM-1-polio virus receptor/nectin-2 pathway [47].

In addition, exosomes isolated from malignant ascites can also impair the cytotoxic activity of peripheral blood mononuclear cells (PBMCs), and Fas ligand (FasL) has shown to be present in the exosomal suspension. Thus, anti-FasL antibody can decrease the percentage of DC and PBMC apoptosis [48]. The Fas/FasL system plays an important role in the immune privilege status of tumors by inducing Fas-mediated apoptosis in tumor-specific lymphocytes. Normal ovarian surface epithelial cells express but do not secret FasL, while EOC cells secrete FasL via the release of exosomes. The release of secreted FasL rather than the membrane-bound form facilitated the escape from immune surveillance and survival of the tumor cells [49]. Lysophosphatidic acid (LPA) can up-regulate FasL presentation on the surface of OC cells and stimulate the secretion of FasL-bearing exosomes, thereby contributing to an OC cell counterattack against activated T cells and promoting OC metastasis [50]. Ovarian carcinoma-derived exosomes have shown to carry phosphatidylserine (PS) on their surface, and PS appeared to be an important molecule for exosomal uptake by NK cells [51]. The PS present on the membrane of exosomes isolated from ovarian tumor tissues or ascites can reversibly inhibit the activation of T cells by arresting the T cell signaling cascade. The depletion or blockade of PS with anti-PS antibody could significantly block the inhibition of T cell function, which represents a potential therapeutic approach for patients with OC [52].

Scientists have leveraged the relation of exosomes to the immune system in antitumor research since 2005. One clinical trial employed an immunotherapeutic approach in conjunction with conventional first-line chemotherapy. The immunotherapy based on the use of a TLR3 agonist and tumor-derived exosomes carrying associated antigens, which could disrupt tumor immune tolerance and immunosuppressive effects. This therapy also exploited tumor-derived exosomes as a potential source of antigens for generating effective and lasting tumor antigen-specific T cell immunity [53]. In another study, metastatic ovarian carcinoma-derived exosomes presented tumor-specific antigens to DCs derived from umbilical cord blood, and these DCs could then stimulate common T cells to differentiate and effectively induce cytotoxicity, thereby providing a novel immunotherapy for OC [54] The potential targets for immunotherapy are listed in Table 2. Although there are currently few applications, the authors believe that the role of exosomes in tumor microenvironment immunity may have the most potential for treating OC.

Table 2 Exosomes provide potential targets for immunotherapy

Exosomes are related to the tumor microenvironment and promote progression and metastasis

OC patients have a poor prognosis and frequently experience recurrence and dissemination, even with chemotherapy or surgical debulking. Recent evidence revealed that the tumor microenvironment consists of stromal cells (including fibroblasts, macrophages, and MDSCs) and extracellular matrix components (including inflammatory cytokines and chemokine), which can promote cancer cell invasion and metastasis [55]. An increasing amount of evidence has shown that exosome-mediated interactions between tumor cells and their microenvironment are a key factor in tumor progression. Exosomes are key mediators of intercellular communication in the tumor microenvironment [56]. Intercellular communication is a function that progressively developed via a process of mutual cellular adaptation. Experiments showed that when cancer cells co-cultured with different individual mesenchymal stem cells (MSCs), the interactions between the MSCs and cancer cells depended on the exchange of exosomes. During these cellular interactions, varieties of functional changes were observed, particularly those promoting the growth of tumor cells [57].

As previously mentioned, urinary exosomal miR-30a-5p could be used as a high-specificity diagnostic marker of OC. At the same time, miR-30a-5p knockdown significantly inhibited OC cell proliferation and migration. Therefore, miR-30a-5p can also serve as a therapeutic target for OSA [26]. It has been found that exosomes, which present in the ascites and blood of ovarian carcinoma patients, contain different exosomal proteins frequently related to tumor progression, such as metalloproteinase inducers (EMMPRIN/CD147) and pro-heparanase. The application of patient-derived exosomes can promote the growth of human SKOV3ip carcinoma xenografts in mice [51]. Although this was just a preliminary animal experiment, the results suggested that the application of exosomes in the treatment of OC was feasible.

Exosome shedding, which is frequently observed in tumor cells, is suggested to be involved in several aspects of tumor progression. It demonstrated that exosomes isolated from invasive tumor cell lines as well as the bodily fluids of OC patients can deliver membrane-type 1 matrix metalloprotease (MT1-MMP), which is involved in matrix degradation and disease progression [58]. Extracellular matrix degradation played an important role in OC metastasis. In OC, ascites-derived exosomes were found to contain gelatin lysis enzymes. In addition, exosome-associated proteolytic activity in the tumor vicinity might augment tumor cell invasion into the stroma. Fractionation of malignant ascites revealed exosomes containing localized extracellular matrix-degrading proteinases. The activation of proteinases, such as MMPs and urokinase-type plasminogen activator (uPA), led to increased extracellular matrix degradation, which may facilitate tumor cell invasion and metastasis [59]. Human OC cell lines (CABA I and A2780) shed exosomes carrying MMPs into the extracellular medium, demonstrating a mechanism for regulating focalized proteolytic activity and interacting with the microenvironment [60]. The L1 adhesion molecule (CD171) overexpressed in EOC and associated with a poor prognosis. Although expressed as a transmembrane molecule, L1 was released from carcinoma cells in a soluble form. Tumor-derived vesicles may be an important source of soluble L1 that can regulate tumor cell function in an autocrine-paracrine fashion. Soluble L1 from ascites showed to be a potent inducer of cell migration [61].

More researches are required to provide insight into the mechanisms underlying exosome-induced invasion and migration. Recent studies revealed that LIN28, an RNA-binding protein, could be highly expressed in the exosomes of OC cells (IGROV1) and be internalized by non-tumor cells. When taken up by HEK293 cells, these exosomes led to significant increases in the expression of genes involved in the epithelial-to-mesenchymal transition (EMT) and induced HEK293 cell invasion and migration [62]. The exosomal release of various miRNAs significantly correlated to the invasive potential of cancer cells, and some miRNAs showed significant associations with effusion site and tumor stage. For example, high miR-21, miR-23b and miR-29a levels were associated with poor progression-free survival, whereas high miR-21 expression was correlated to poor overall survival [63]. OC cells specifically secrete exosomes with miRNAs. Then, these exosomes communicate with mesothelial cells to allow tumor dissemination through the peritoneal cavity. When comparing OC cell lines with high (Skov-3) and low (OVCaR-3) invasive potential, transcripts of the miR-let-7 family were more abundant in OVCaR-3 cells than Skov-3 cells. In contrast, they were more abundant in exosomes from Skov-3 cells than OVCaR-3 cells. Transcripts from the miR-200 family was only identified in OVCaR-3 cells and their exosomes [64]. These findings might because miR-let-7 suppresses cell proliferation by transporting more exosomes outside of cells, while miR-200 suppresses the EMT.

Malignant ascites is a natural medium for cell communication. A cluster of spliceosomal proteins identified in ascites as prognostic markers for the prediction of patient survival. Furthermore, they played a role in communication between cancer cells [30]. As previously mentioned, the pentose phosphate pathway is a dominant mechanism mediating intracellular communication via exosomes excreted from OC cells. Therefore, G6PD, TK, and TA1 are key proteins that could become therapeutic targets of OC [40].

A novel concept for the treatment of OC based on impairing the crosstalk between metastatic ovarian cells and their environment has been tested. And a tumor cell capture device was fabricated with exosomes and a three-dimensional scaffold (metastatic trap [M-Trap]). M-Trap transformed a systemic disease into a focalized disease, in which proven therapeutic approaches, such as surgery, can extend survival [65]. Certainly, there are exosomes that directly possess antitumor effects, such as disintegrin and metalloproteinase 15 (ADAM15)-rich exosomes. ADAM15 can be released into the extracellular space as a component cleaved from released exosomes, and this effectively inhibited cancer cell migration and activation [66]. In another study, exosomes from MSCs inhibited cell cycle progression in all tested cell lines and induced Skov-3 cell necrosis. An increase in the number of cells in the G0-G1 phase was observed, suggesting a block in cell cycle progression. These results indicated that exosomes could halt proliferation, thus leading to cell death by apoptosis or necrosis [67]. The functions of exosomes described previously are showed in Table 3.

Table 3 Exosomes related to tumor progression, metastasis and angiogenesis

Exosomes are related to angiogenesis

To date, curative therapies for OC remain under development, and anti-angiogenesis approaches may provide alternative strategies. Tumor-derived exosomes are essential for tumor angiogenesis.

High-grade OC-derived exosomes have a more profound impact on angiogenesis than those derived from low-grade OC. Proteomic profiles revealed that the constituents of high-grade OC-derived exosomes, such as activating transcription factor 2 (ATF2) and metastasis-associated protein 1 (MTA1), among others, had a profound impact on angiogenesis and may play a key role in enhancing tumor development [27]. CD147 is a membrane molecule that is highly expressed in tumor cells and is involved in the progression of malignancies by regulating MMP expression in peritumoral stromal cells. A recent study showed that CD147 expressed in exosomes derived from EOC cells and that CD147-positive vesicles shed by OC cells may induce angiogenesis activities in human umbilical vein endothelial cells [68].

Anti-angiogenesis effects may serve as the mechanisms of some chemotherapy drugs and be expected to provide targets for the treatment of OC. Recently, a study demonstrated the anti-neoplastic effect of Amla extract (Emblica officinalis, AE) on OC in vitro and in vivo. The underlying mechanism was that AE could down-regulate pro-angiogenic molecules, such as insulin-like growth factor 1 receptor (IGF1R), via up-regulating cellular and exosomal miR-375 in human OC cells [69], explaining why AE attenuated angiogenesis and had an antitumor effect.

Exosomes are related to drug resisitance

There is no doubt that OC is the most lethal gynecological malignancy. Most OC patients are initially sensitively responsive to the preferred chemotherapy with platinum and paclitaxel (PTX). Unfortunately, most patients experience recurrence within 6–12 months and die of progressively chemotherapy-resistant disease [70].

One study showed that cancer-associated adipocytes (CAAs) or fibroblasts (CAFs) can secrete exosomes to transfer miR-21 into neighboring cancer cells, thereby increasing the chemo-resistance of these cells to PTX via the down-regulation of the direct target apoptotic protease-activating factor 1 (APAF1). Exosomal miR-21 can enhance OC cell PTX-resistance and decrease OC cell apoptosis [71]. The relationship between endogenous miR-433 expression and resistance to PTX was also been investigated in OC cell lines. Moreover, miR-433 mediated the down-regulation of cyclin-dependent kinase 6 (CDK6) [72]. Interestingly, miR-21-3p induced cisplatin (CIS) resistance in ovarian tumors, potentially by targeting the neuron navigator 3 (NAV3) gene. Exosomes released by CIS-resistant cells were also capable of increasing resistance in other cells [73].

In addition to miRNAs, exosomal proteins are also involved in the process of tumor drug resistance. After being shed, exosomes can transfer their contents to recipient cells at the membrane-bound protein level. For example, P-glycoprotein (P-gp) shared between human OC cells via intercellular exosomal transfer, resulting in enhanced cellular resistance to PTX [74]. Jie Y et al. found that increased annexin A3 expression was a mechanism for platinum resistance in OC cells [75]. Further research showed that serum annexin A3 levels were significantly higher in platinum-resistant patients than in platinum-sensitive patients. It was also found that annexin A3 was associated with exosomal release from platinum-resistant OC cells, which was supported by OC cells expressing higher levels of annexin A3 releasing increased numbers of exosomes [76]. Proteins may be transports or carriers of chemotherapy drugs. The lysosomal proteins ATPase copper-transporting alpha and beta (ATP7A and ATP7B), which are putative CIS-export transporters, can enhance the exosomal export of platinum in drug-resistant human ovarian carcinoma cells [77].

Now, the supporting role on drug resistance of exosomes sourced from tumor microenvironment has, gain more and more attention, such as exosomes from CAAs or CAFs. Although there are few related research in ovarian cancer, we believe that they will increase rapidly, thus we still listed them separately in Table 4.

Table 4 Exosomes related to drug resistance

Conclusions

Due to the asymptomatic nature of this disease during the early stages and resistance to chemotherapy, OC, particularly EOC, has a poor prognosis [5]. The distinct components and biological activities of exosomes must be explored to identify novel high-sensitivity and high-specificity biomarkers for the early diagnosis of EOC. Such exosomal biomarkers should significantly increase in OC patients compared with healthy individuals, which are higher in patients with advanced-stage disease than early-stage disease. Some miRNAs meet these requirements, such as miR-222-3p [22], miR-200b and miR-200c [24]. However, as previously mentioned, related therapies remain in the laboratory stage and there is a lack of large-scale clinical trials. In addition, the diagnostic potential of any biomarker depends on the sample type (e.g., serum and urine, among others), and the maturity of the purification and detection methods will be a constraint. The limitations of current clinical OC therapies require the development of new treatments. Cancer-related exosomes exhibit powerful influences on tumors by a variety of biological mechanisms, and it is important to understand the underlying mechanisms of exosome-mediated immunosuppression and tumor progression, metastasis and angiogenesis. The ability of exosomes block factors in tumor microenvironments should enhance antitumor responses in OC. Most EOC patients died of progressively chemotherapy-resistant disease notably [70]. In addition to miRNAs, exosomal proteins are also involved in the process of drug resistance in OC patients. In order to retain patient sensitivity to platinum and PTX, the mechanism of drug resistance should be more thoroughly understood, and exosomes provide a novel perspective in that context.

In summary, the emerging field of exosomes has created new perspectives on diagnosis, prognosis, treatment, and drug resistance in OC (Fig. 1). Exosomes represent a mobile source of information regarding the molecular makeup of parental tumors. However, the exosomal knowledgebase has thus far been limited to the findings of a few persuasive studies. In addition, high-throughput exosomal isolation and analysis remains difficult. Much more research on exosomes needs to be perform to reveal breakthroughs in diagnosing and treating OC.

Fig. 1
figure 1

A summary of the roles of exosomes in ovarian cancer

Abbreviations

ADAM:

a disintegrin and metalloproteinase

ALCAM:

Activated leukocyte cell adhesion molecule

APAF1:

Apoptotic protease-activating factor 1

ATF2:

Activating transcription factor 2

ATP7A:

ATPase copper-transporting alpha

ATP7B:

ATPase copper-transporting beta

CAAs:

Cancer-associated adipocytes

CAFs:

Cancer-associated fibroblasts

CC:

Colon carcinoma

CDK6:

Cyclin-dependent kinase 6

CIS:

Cisplatin

DCs:

Dendritic cells

DNAM-1:

CD226

EMMPRIN:

Metalloproteinase-inducer

EMT:

Epithelial-to-mesenchymal transition

EpCAM:

Epithelial cell adhesion molecule

FasL:

Fas ligand

G6PD:

Glucose-6-phosphate dehydrogenase

GC:

Gastric cancer

HIF-1α:

Hypoxia-inducible factor-1 alpha

IGF1R:

Insulin-like growth factor 1 receptor

LPA:

Lysophosphatidic acid

M2:

Macrophage 2 phenotypes

MDSCs:

Myeloid-derived suppressor cells

MSCs:

Mesenchymal stem cells

MT1-MMP:

Membrane-type 1 matrix metalloprotease

MTA1:

Metastasis-associated protein 1

NAV3:

Neuron navigator 3

NK:

Natural killer

NKG2D:

NK C-type lectin-like receptor

OC:

Ovarian cancer

OSA:

Ovarian serous adenocarcinoma

OSC:

Ovarian serous carcinoma

PBMCs:

Peripheral blood mononuclear cells

P-gp:

P-glycoprotein

PS:

Phosphatidylserine

PTX:

Paclitaxel

TA1:

Transaldolase 1

TK:

Transketolase

TLR:

Toll-like receptor

uPA:

urokinase-type plasminogen activator

VAMP3:

Vesicle-associated membrane protein 3

VEGF:

Vascular endothelial growth factor

WT:

Wild type

References

  1. Wanqing C, Rongshou Z, Baade PD. Cancer statistics in China, 2015. CA Cancer J Clin. 2016;66:115–32.

    Article  Google Scholar 

  2. Siegel LR, Miller KD, Jemal A, DVM. Cancer statistics, 2015. CA Cancer J Clin. 2015;65:5–29.

    Article  PubMed  Google Scholar 

  3. Stack MS, Fishman DA. Ovarian Cancer, Cancer Treatment and Research, vol. 149. US: Springer; 2010.

    Book  Google Scholar 

  4. Cancer Research UK. Cancer stats report–ovarian cancer. 2011. accessed July 3, 2014.

  5. Coleman RL, Monk BJ, Sood AK, Herzog TJ. Latest research and treatment of advanced-stage epithelial ovarian cancer. Nat Rev Clin Oncol. 2013;10:211–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Johnstone RM, Adam M, Hammond JR, Orr L, Turbide C. Vesicle formation during reticulocyte maturation. Association of plasma membrane activities with released vesicles (exosomes). J Biol Chem. 1987;262:9412–20.

    CAS  PubMed  Google Scholar 

  7. Raposo G, Stoorvogel W. Extracellular vesicles: exosomes, microvesicles, and friends. J Cell Biol. 2013;200:373–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Giusti I, D’Ascenzo S, Dolo V. Microvesicles as potential ovarian cancer biomarkers. Biomed Res Int. 2013;2013:703048.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Lara M, Amit S, George M, Megha S, Amiji MM. Exosome mediated communication within the tumor microenvironment. J Control Release. 2015;219:278–94.

    Article  Google Scholar 

  10. Hong BS, Cho JH, Kim H, et al. Colorectal cancer cell-derived microvesicles are enriched in cell cycle-related mRNAs that promote proliferation of endothelial cells. BMC Genom. 2009;10:556.

    Article  Google Scholar 

  11. Kogure T, Lin WL, Yan IK, Braconi C, Patel T. Intercellular nanovesicle-mediated microRNA transfer: a mechanism of environmental modulation of hepatocellular cancer cell growth. Hepatology. 2011;54:1237–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Zhang L, Zhang S, Yao J, et al. Microenvironment-induced PTEN loss by exosomal microRNA primes brain metastasis outgrowth. Nature. 2015;527:100–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Challagundla KB, Wise PM, Neviani P, et al. Exosome‐mediated transfer of microRNAs within the tumor microenvironment and neuroblastoma resistance to chemotherapy. J Natl Cancer Inst. 2015;107:djv135.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Kosaka N, Iguchi H, Hagiwara K, Yoshioka Y, Takeshita F, Ochiya T. Neutral sphingomyelinase 2 (nSMase2) -dependent exosomal transfer of angiogenic microRNAs regulate cancer cell metastasis. J Biol Chem. 2013;288:10849–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Fabbri M, Paone A, Calore F, et al. MicroRNAs bind to Toll-like receptors to induce prometastatic inflammatory response. Proc Natl Acad Sci U S A. 2012;109:E2110–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Fong MY, Zhou W, Liu L, et al. Breast-cancer-secreted miR-122 reprograms glucose metabolism in premetastatic niche to promote metastasis. Nat Cell Biol. 2015;17:183–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Tang K, Zhang Y, Zhang H, Xu P, Liu J, Ma J, Lv M. Delivery of chemotherapeutic drugs in tumour cell-derived microparticles. Nat Commun. 2012;3:1282.

    Article  PubMed  Google Scholar 

  18. Graves LE, Ariztia EV, Navari JR, et al. Proinvasive properties of ovarian cancer ascites-derived membrane vesicles. Cancer Res. 2004;64:7045–9.

    Article  CAS  PubMed  Google Scholar 

  19. Kajdos M, Janas L, Kolasa-Zwierzchowska D, Wilczyński JR, Stetkiewicz T. Microvesicles as a potential biomarker of neoplastic diseases and their role in development and progression of neoplasm. Prz Menopauzalny. 2015;14(4):283–91.

    PubMed  PubMed Central  Google Scholar 

  20. Mikaelian I, Scicchitano M, Mendes O, et al. Frontiers in preclinical safety biomarkers: microRNAs and messenger RNAs. Toxicol Pathol. 2013;41:18–31.

    Article  CAS  PubMed  Google Scholar 

  21. Taylor DD, Gercel-Taylor C. MicroRNA signatures of tumor-derived exosomes as diagnostic biomarkers of ovarian cancer. Gynecol Oncol. 2008;110(1):13–21.

    Article  CAS  PubMed  Google Scholar 

  22. Xiang Y, Quanfeng W, Xiaoli W, Qinyi Z, Xinjing W, Lu J, Xin C. Epithelial ovarian cancer-secreted exosomal miR-222-3p induces polarization of tumor-associated macrophages. Oncotarget. 2016;7(28):43076–87.

  23. Cappellesso R, Tinazzi A, Giurici T, Simonato F, Guzzardo V, Ventura L, Crescenzi M. Programmed cell death 4 and microRNA 21 inverse expression is maintained in cells and exosomes from ovarian serous carcinoma effusions. Cancer Cytopathol. 2014;122(9):685–93.

    Article  CAS  PubMed  Google Scholar 

  24. Meng X, Müller V, Milde-Langosch K, Trillsch F, Pantel K, Schwarzenbach H. Diagnostic and prognostic relevance of circulating exosomal miR-373, miR-200a, miR-200b and miR-200c in patients with epithelial ovarian cancer. Oncotarget. 2016;7(13):16923–35.

  25. Zavesky L, Jandakova E, Turyna R, Langmeierova L, Weinberger V, Minar L. Supernatant versus exosomal urinary microRNAs. Two fractions with different outcomes in gynaecological cancers. Neoplasma. 2016;63(1):121–32.

    Article  CAS  PubMed  Google Scholar 

  26. Zhou J, Gong G, Tan H, Dai F, Zhu X, Chen Y, Wang J. Urinary microRNA-30a-5p is a potential biomarker for ovarian serous adenocarcinoma. Oncol Rep. 2015;33(6):2915–23.

    CAS  PubMed  Google Scholar 

  27. Yi H, Ye J, Yang XM, Zhang LW, Zhang ZG, Chen YP. High-grade ovarian cancer secreting effective exosomes in tumor angiogenesis. Int J Clin Exp Pathol. 2015;8:5062–70.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Sinha A, Ignatchenko V, Ignatchenko A, Mejia-Guerrero S, Kislinger T. In-depth proteomic analyses of ovarian cancer cell line exosomes reveal differential enrichment of functional categories compared to the NCI 60 proteome. Biochem Biophys Res Commun. 2014;445(4):694–701.

    Article  CAS  PubMed  Google Scholar 

  29. Liang B, Peng P, Chen S, Li L, Zhang M, Cao D, Yang J. Characterization and proteomic analysis of ovarian cancer-derived exosomes. J Proteomics. 2013;80:171–82.

    Article  CAS  PubMed  Google Scholar 

  30. Shender VO, Pavlyukov MS, Ziganshin RH, Arapidi GP, Kovalchuk SI, Anikanov NA, Altukhov IA. Proteome-metabolome profiling of ovarian cancer ascites reveals novel components involved in intercellular communication. Mol Cell Proteomics. 2014;13(12):3558–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Szajnik M, Derbis M, Lach M, Patalas P, Michalak M, Drzewiecka H, Szpurek D. Exosomes in Plasma of Patients with Ovarian Carcinoma: Potential Biomarkers of Tumor Progression and Response to Therapy. Gynecol Obstet (Sunnyvale). 2013;Suppl 4:3.

  32. Fan Q, Cheng JC, Qiu X, Chang HM, Leung PC. EpCAM is up-regulated by EGF via ERK1/2 signaling and suppresses human epithelial ovarian cancer cell migration. Biochem Biophys Res Commun. 2015;457(3):256–61.

    Article  CAS  PubMed  Google Scholar 

  33. Runz S, Keller S, Rupp C, Stoeck A, Issa Y, Koensgen D, Mustea A. Malignant ascites-derived exosomes of ovarian carcinoma patients contain CD24 and EpCAM. Gynecol Oncol. 2007;107(3):563–71.

    Article  CAS  PubMed  Google Scholar 

  34. Beach A, Zhang HG, Ratajczak MZ, Kakar SS. Exosomes: an overview of biogenesis, composition and role in ovarian cancer. J Ovarian Res. 2014;7:14.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Speicher MR, Pantel K. Tumor signatures in the blood. Nat Biotechnol. 2014;32(5):441–3.

    Article  CAS  PubMed  Google Scholar 

  36. Im H, Shao H, Yong Il P, Peterson VM, Castro CM, Ralph W. Label-free detection and molecular profiling of exosomes with a nano-plasmonic sensor. Nat Biotechnol. 2014;32:488–93.

    Article  Google Scholar 

  37. Carbotti G, Orengo AM, Mezzanzanica D, Bagnoli M, Brizzolara A, Emionite L, Puppo A. Activated leukocyte cell adhesion molecule soluble form: a potential biomarker of epithelial ovarian cancer is increased in type II tumors. Int J Cancer. 2013;132(11):2597–605.

    Article  CAS  PubMed  Google Scholar 

  38. Taylor DD, Gercel-Taylor C, Parker LP. Patient-derived tumor-reactive antibodies as diagnostic markers for ovarian cancer. Gynecol Oncol. 2009;115(1):112–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Li J, Sherman-Baust CA, Tsai-Turton M, Bristow RE, Roden RB, Morin PJ. Claudin-containing exosomes in the peripheral circulation of women with ovarian cancer. BMC Cancer. 2009;9:244.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Yi H, Zheng X, Song J, Shen R, Su Y, Lin D. Exosomes mediated pentose phosphate pathway in ovarian cancer metastasis: a proteomics analysis. Int J Clin Exp Pathol. 2015;8(12):15719–28.

    PubMed  PubMed Central  Google Scholar 

  41. Gomes J, Gomes-Alves P, Carvalho SB, Peixoto C, Alves PM, Altevogt P. Costa JExtracellular Vesicles from Ovarian Carcinoma Cells Display Specific Glycosignatures. Biomolecules. 2015;5(3):1741–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Zhao Z, Yang Y, Zeng Y, He M. A microfluidic ExoSearch chip for multiplexed exosome detection towards blood-based ovarian cancer diagnosis. Lab Chip. 2016;16(3):489–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Filipazzi P, Bürdek M, Villa A, Rivoltini L, Huber V. Recent advances on the role of tumor exosomes in immunosuppression and disease progression. Semin Cancer Biol. 2012;22:342–9.

    Article  CAS  PubMed  Google Scholar 

  44. Kumar V, Patel S, Tcyganov E, Gabrilovich DI. The nature of myeloid-derived suppressor cells in the tumor microenvironment. Trends Immunol. 2016;37(3):208–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Taylor DD, Gerçel-Taylor C. Tumour-derived exosomes and their role in cancer-associated T-cell signalling defects. Br J Cancer. 2005;92(2):305–11.

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Bretz NP, Ridinger J, Rupp AK, Rimbach K, Keller S, Rupp C, Marmé F. Body fluid exosomes promote secretion of inflammatory cytokines in monocytic cells via Toll-like receptor signaling. J Biol Chem. 2013;288(51):36691–702.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Labani-Motlagh A, Israelsson P, Ottander U, Lundin E, Nagaev I, Nagaeva O, Dehlin E. Differential expression of ligands for NKG2D and DNAM-1 receptors by epithelial ovarian cancer-derived exosomes and its influence on NK cell cytotoxicity. Tumour Biol. 2016;37(4):5455–66.

    Article  CAS  PubMed  Google Scholar 

  48. Peng P, Yan Y, Keng S. Exosomes in the ascites of ovarian cancer patients: origin and effects on anti-tumor immunity. Oncol Rep. 2011;25(3):749–62.

    CAS  PubMed  Google Scholar 

  49. Abrahams VM, Straszewski SL, Kamsteeg M, Hanczaruk B, Schwartz PE, Rutherford TJ, Mor G. Epithelial ovarian cancer cells secrete functional Fas ligand. Cancer Res. 2003;63(17):5573–81.

    CAS  PubMed  Google Scholar 

  50. Meng Y, Kang S, Fishman DA. Lysophosphatidic acid stimulates fas ligand microvesicle release from ovarian cancer cells. Cancer Immunol Immunother. 2005;54(8):807–14.

    Article  CAS  PubMed  Google Scholar 

  51. Keller S, König AK, Marmé F, Runz S, Wolterink S, Koensgen D, Mustea A. Systemic presence and tumor-growth promoting effect of ovarian carcinoma released exosomes. Cancer Lett. 2009;278(1):73–81.

    Article  CAS  PubMed  Google Scholar 

  52. Kelleher Jr RJ, Balu-Iyer S, Loyall J, Sacca AJ, Shenoy GN, Peng P, Iyer V. Extracellular vesicles present in human ovarian tumor microenvironments induce a phosphatidylserine-dependent arrest in the T-cell signaling cascade. Cancer Immunol Res. 2015;3(11):1269–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Adams M, Navabi H, Croston D, Coleman S, Tabi Z, Clayton A, Jasani B. The rationale for combined chemo/immunotherapy using a Toll-like receptor 3 (TLR3) agonist and tumour-derived exosomes in advanced ovarian cancer. Vaccine. 2005;23(17–18):2374–8.

    CAS  PubMed  Google Scholar 

  54. Li QL, Bu N, Yu YC, Hua W, Xin XY. Exvivo experiments of human ovarian cancer ascites-derived exosomes presented by dendritic cells derived from umbilical cord blood for immunotherapy treatment. Clin Med Oncol. 2008;2:461–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Luo Z, Wang Q, Lau WB, Lau B, Xu L, Zhao L, Yang H. Tumor microenvironment: the culprit for ovarian cancer metastasis? Cancer Lett. 2016;377(2):174–82.

    Article  CAS  PubMed  Google Scholar 

  56. Tang MK, Wong AS. Exosomes: emerging biomarkers and targets for ovarian cancer. Cancer Lett. 2015;367(1):26–33.

    Article  CAS  PubMed  Google Scholar 

  57. Yang Y, Otte A, Hass R. Human mesenchymal stroma/stem cells exchange membrane proteins and alter functionality during interaction with different tumor cell lines. Stem Cells Dev. 2015;24(10):1205–22.

    Article  CAS  PubMed  Google Scholar 

  58. Clancy JW, Sedgwick A, Rosse C, Muralidharan-Chari V, Raposo G, Method M, Chavrier P. Regulated delivery of molecular cargo to invasive tumour-derived microvesicles. Nat Commun. 2015;6:6919.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Graves LE, Ariztia EV, Navari JR, Matzel HJ, Stack MS, Fishman DA. Proinvasive properties of ovarian cancer ascites-derived membrane vesicles. Cancer Res. 2004;64(19):7045–9.

    Article  CAS  PubMed  Google Scholar 

  60. Dolo V, D’Ascenzo S, Giusti I, Millimaggi D, Taraboletti G, Pavan A. Shedding of membrane vesicles by tumor and endothelial cells. Ital J Anat Embryol. 2005;110(2 Suppl 1):127–33.

    CAS  PubMed  Google Scholar 

  61. Gutwein P, Stoeck A, Riedle S, Gast D, Runz S, Condon TP, Marmé A. Cleavage of L1 in exosomes and apoptotic membrane vesicles released from ovarian carcinoma cells. Clin Cancer Res. 2005;11(7):2492–501.

    Article  CAS  PubMed  Google Scholar 

  62. Enriquez VA, Cleys ER, Da Silveira JC, Spillman MA, Winger QA, Bouma GJ. High LIN28A expressing ovarian cancer cells secrete exosomes that induce invasion and migration in HEK293 cells. Biomed Res Int. 2015;2015:701390.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Vaksman O, Tropé C, Davidson B, Reich R. Exosome-derived miRNAs and ovarian carcinoma progression. Carcinogenesis. 2014;35(9):2113–20.

    Article  CAS  PubMed  Google Scholar 

  64. Kobayashi M, Salomon C, Tapia J, Illanes SE, Mitchell MD, Rice GE. Ovarian cancer cell invasiveness is associated with discordant exosomal sequestration of Let-7 miRNA and miR-200. J Transl Med. 2014;12:4.

    Article  PubMed  PubMed Central  Google Scholar 

  65. de la Fuente A, Alonso-Alconada L, Costa C, Cueva J, Garcia-Caballero T, Lopez-Lopez R, Abal M. M-trap: exosome-based capture of tumor cells as a new technology in peritoneal metastasis. J Natl Cancer Inst. 2015;6:107(9).

    Google Scholar 

  66. Lee HD, Kim YH, Koo BH, Kim DS. The ADAM15 ectodomain is shed from secretory exosomes. BMB Rep. 2015;48(5):277–82.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Bruno S, Collino F, Deregibus MC, Grange C, Tetta C, Camussi G. Microvesicles derived from human bone marrow mesenchymal stem cells inhibit tumor growth. Stem Cells Dev. 2013;22(5):758–71.

    Article  CAS  PubMed  Google Scholar 

  68. Millimaggi D, Mari M, D’Ascenzo S, Carosa E, Jannini EA, Zucker S, Carta G. Tumor vesicle-associated CD147 modulates the angiogenic capability of endothelial cells. Neoplasia. 2007;9(4):349–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. De A, Powers B, De A, Zhou J, Sharma S, Van Veldhuizen P, Bansal A. Emblica officinalis extract downregulates pro-angiogenic molecules via upregulation of cellular and exosomal miR-375 in human ovarian cancer cells. Oncotarget. 2016. doi: 10.18632/oncotarget.8966.

  70. Cho KR, Shih IM. Ovarian cancer. Annu Rev Pathol. 2009;4:287–313.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Au Yeung CL, Co NN, Tsuruga T, Yeung TL, Kwan SY, Leung CS, Li Y. Exosomal transfer of stroma-derived miR21 confers paclitaxel resistance in ovarian cancer cells through targeting APAF1. Nat Commun. 2016;7:11150.

    Article  PubMed  PubMed Central  Google Scholar 

  72. Weiner-Gorzel K, Dempsey E, Milewska M, McGoldrick A, Toh V, Walsh A, Lindsay S. Overexpression of the microRNA miR-433 promotes resistance to paclitaxel through the induction of cellular senescence in ovarian cancer cells. Cancer Med. 2015;4(5):745–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Pink RC, Samuel P, Massa D, Caley DP, Brooks SA, Carter DR. The passenger strand, miR-21-3p, plays a role in mediating cisplatin resistance in ovarian cancer cells. Gynecol Oncol. 2015;137(1):143–51.

    Article  CAS  PubMed  Google Scholar 

  74. Zhang FF, Zhu YF, Zhao QN, Yang DT, Dong YP, Jiang L, Xing WX. Microvesicles mediate transfer of P-glycoprotein to paclitaxel-sensitive A2780 human ovarian cancer cells, conferring paclitaxel-resistance. Eur J Pharmacol. 2014;738:83–90.

    Article  CAS  PubMed  Google Scholar 

  75. Yan XD, Yin J, Yao H, et al. Increased expression of annexin A3 is a mechanism of platinum resistance in ovarian cancer. Cancer Res. 2010;70:1616–24.

    Article  CAS  PubMed  Google Scholar 

  76. Yin J, Yan X, Yao X, Zhang Y, Shan Y, Mao N, Yang Y. Secretion of annexin A3 from ovarian cancer cells and its association with platinum resistance in ovarian cancer patients. J Cell Mol Med. 2012;16(2):337–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Safaei R, Larson BJ, Cheng TC, Gibson MA, Otani S, Naerdemann W, Howell SB. Abnormal lysosomal trafficking and enhanced exosomal export of cisplatin in drug-resistant human ovarian carcinoma cells. Mol Cancer Ther. 2005;4(10):1595–604.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

We thank the board of the Shanghai First Maternity and Infant Hospital Corporation for their excellent technical assistance.

Funding

This study was supported by grants from the National Natural Science Foundation of China (81372787, 81072136), Shanghai Municipal Bureau of Health (20134033) Shanghai Health System joint research project (2013ZYJB0201) and Top 100 Medical Elite in Shanghai (XBR 2011065).

Availability of data and materials

Not applicable.

Authors’ contributions

XDL designed the figures, researched appropriate references and was a major contributor in writing the manuscript. XPW developed the structure of the article and guided the selection of references. Both authors read and approved the final manuscript.

Competing interests

The authors declare that they have no competing interests.

Consent for publication

Both authors have approved the final manuscript for publication.

Ethics approval and consent to participate

Not applicable.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xipeng Wang.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, X., Wang, X. The emerging roles and therapeutic potential of exosomes in epithelial ovarian cancer. Mol Cancer 16, 92 (2017). https://doi.org/10.1186/s12943-017-0659-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12943-017-0659-y

Keywords