Next Article in Journal
Tunable Drug Release Rate Using Modular Oral Dosage Forms
Previous Article in Journal
Membrane Vesicles as Drug Delivery Systems: Source, Preparation, Modification, Drug Loading, In Vivo Administration and Biodistribution, and Application in Various Diseases
Previous Article in Special Issue
Exosomes: Diagnostic and Therapeutic Implications in Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

New Possible Ways to Use Exosomes in Diagnostics and Therapy via JAK/STAT Pathways

1
Department of Human Genetics, Faculty of Medicine, University of Debrecen, Egyetem Tér 1, H-4032 Debrecen, Hungary
2
Comenius University Science Park, 841 04 Bratislava, Slovakia
3
Geneton Ltd., 841 04 Bratislava, Slovakia
4
Institute of Medical Biology, Genetics and Clinical Genetics, Faculty of Medicine, Comenius University, 811 08 Bratislava, Slovakia
5
Department of Molecular Biology, Faculty of Natural Sciences, Comenius University, 841 01 Bratislava, Slovakia
6
3rd Surgical Clinic, Faculty of Medicine, Comenius University and Merciful Brothers University Hospital, 811 08 Bratislava, Slovakia
7
Division of Clinical Genetics, Department of Laboratory Medicine, Faculty of Medicine, University of Debrecen, H-4032 Debrecen, Hungary
*
Author to whom correspondence should be addressed.
Pharmaceutics 2023, 15(7), 1904; https://doi.org/10.3390/pharmaceutics15071904
Submission received: 24 May 2023 / Revised: 27 June 2023 / Accepted: 4 July 2023 / Published: 7 July 2023
(This article belongs to the Special Issue Exosomes as Drug Carriers for Cancer Therapy)

Abstract

:
Exosomes have the potential to be the future of personalized diagnostics and therapy. They are nano-sized particles between 30 and 100 nm flowing in the extracellular milieu, where they mediate cell–cell communication and participate in immune system regulation. Tumor-derived exosomes (TDEs) secreted from different types of cancer cells are the key regulators of the tumor microenvironment. With their immune suppressive cargo, TDEs prevent the antitumor immune response, leading to reduced effectiveness of cancer treatment by promoting a pro-tumorigenic microenvironment. Involved signaling pathways take part in the regulation of tumor proliferation, differentiation, apoptosis, and angiogenesis. Signal transducers and activators of transcription factors (STATs) and Janus kinase (JAK) signaling pathways are crucial in malignancies and autoimmune diseases alike, and their potential to be manipulated is currently the focus of interest. In this review, we aim to discuss exosomes, TDEs, and the JAK/STAT pathways, along with mediators like interleukins, tripartite motif proteins, and interferons.

1. Introduction

Exosomes are nano-sized extracellular vesicles carrying donor cell-derived markers such as nucleic acids, proteins (including receptors and enzymes), and lipids that may provide practical implications relying on the fact that most cell types secrete exosomes easily accessible by non-invasive liquid biopsy-based approaches [1]. These vesicles can mediate cancer progression and metastasis or reduce the cytotoxic effect of antitumor therapy [2]. On the other hand, exosomes can be engineered and delivered to recipient cells to modify the tumor microenvironment and enhance the effectiveness of cancer therapy in patients [3]. Nowadays JAK/STAT is also procured in the cancer-associated pathways, thus its modulation via exosomes is gaining attention in potential therapeutic applications.

2. Exosome Biogenesis

Exosomes are unique vesicles carrying numerous biomarkers, that may provide potential for personalized diagnostics and therapy. They are nano-sized particles between 30 and 100 nm in the extracellular milieu, mediating cell–cell communication and participating in the immune system. The exosome biogenesis starts with the intrusion of the plasma membrane, leading to an early endosomal shape in which low-density vesicles are very close to the surface of the inner membrane [4,5]. A particular collection of cytosolic factors on the endosomal membrane manages the invagination of the endosomal lipid bilayer towards their luminal space, leading to the creation of intraluminal vesicles. Such a process, called early endosome maturation, results in the formation of multivesicular bodies [6] that can fuse with cell membranes via an adenosine triphosphate (ATP)-dependent process, thus releasing exosomes into the neighboring extracellular space [7]. In terms of structure, exosomes are encompassed by a phospholipid bilayer containing integral membrane proteins and lipid rafts. Some transmembrane proteins are found in a glycosylated form, which allows other proteins to bind using their carbohydrate side chains. The role of the outer bilayer is to protect the molecules transported inside, while the content is regulated by the endosomal sorting complex required for transport [4,5]. Among the transported molecules, annexins, flotillins, Rab proteins, CD8, CD13, CD14, CD9, CD63, CD81, and many other proteins characteristic of the donor cell may be present. In addition to proteins, exosomes may also contain different nucleic acids, e.g., DNA, mitochondrial DNA (mtDNA), RNAs, such as microRNA (miRNA), long non-coding RNA (lncRNA), and piwi-interacting RNA (piRNA) (Figure 1) [5,8,9,10,11,12,13,14,15].

2.1. Exosomal Cargo Uptake via JAK/STAT Dependent Pathway

Recipient cells may incorporate exosomal cargo through membrane diffusion or via a receptor-mediated pathway employing tripartite motif (TRIM) proteins [5,8,9,10,11,12,13,14]. The TRIM protein family is known for its regulatory roles in transcription, apoptosis, and cellular growth, and their effect on the Janus kinases (JAKs) and signal transducer and the activator of transcription proteins (STATs) signaling pathways has also been described [16]. Some TRIMs are frequently found in specific tumors. For example, TRIM44 has modulatory effects on tumor progression and is a potential indicator of unfavorable prognosis in epithelial ovarian cancer [17]. TRIM44 expression is positively regulated via ovarian cancer-derived exosomal circRNA nuclear factor I X (circNFIX) [18]. The JAK/STAT pathway is regulated by ovarian cancer-derived exosomal circNFIX in human umbilical vein endothelial cells through the miR518a-3p/TRIM44 axis (Table 1). circNFIX and TRIM44 correlate with the tumor stage as defined by the International Federation of Gynecology and Obstetrics (FIGO) and with the size of ovarian tumors [18]. The effect of TRIM44 was shown to be modified in a microRNA-dependent manner. The miR-34a-5p may directly target TRIM44, thus inhibiting its activity and leading to suppression of ovarian cancer angiogenesis and malignant behavior [18]. Also, TRIM66 may advance the expression of the JAK/STAT pathways in prostate carcinoma. TRIM66 is part of the signaling axis composed of TRIM66-STAT2-IL-2 molecules, in which TRIM66 favorably regulates the expression of STAT2 and IL-2. It was found that overexpression of either STAT2 or IL-2 almost completely abolished the inhibitory effects of TRIM66 deficiency on cell proliferation, migration, and invasion. This demonstrates the predominance of STAT2-IL-2 in mediating the oncogenic properties of TRIM66 [19].
Table 1. The role of TDE-derived RNAs, DNA, and proteins in onco-immunology and JAK/STAT pathways in various cancer types.
Table 1. The role of TDE-derived RNAs, DNA, and proteins in onco-immunology and JAK/STAT pathways in various cancer types.
CancerSourceBiomarkerFunctionReferences
Ovarian cancerRNAcircNFIX, miR-518a-3pCorrelates with FIGO and tumor size progression[18]
Lung cancerRNAmiR-21, miR-29aConnects to TLR-8 pro-inflammatory cytokines, metastasis, tumor proliferation[2]
Multiple myelomaRNAmiR-146aActivates JAK-STAT pathways via Notch pathway[3,20,21,22]
Gastric cancerRNAmiR-3184-5pSuppresses cell proliferation, migration, and invasion downregulating inter alia the expression of p-STAT3[23,24,25]
Prostate cancerRNAmiR-187Suppresses malignancy via JAK3/STAT3 pathway[26]
RNAmiR-222-3pOverexpression promotes cell proliferation[27]
Renal cell carcinomaRNAlncARSRinducing macrophage polarization by activating STAT3[28]
NSCLCDNAmutated KRASSwitch naive CD4+ T cells to Treg-like cells[29]
Pancreatic cancerProteinKRASPromotes macrophage polarization (M2) via STAT3[30]
Breast cancerProteingp130Increasing the levels of pro-tumorigenic cytokines via STAT3[31]
Lung cancerProteinhsp72STAT3-dependent immunosuppressive function[2]

2.2. Biological Roles of Tumor-Derived Exosomes in Tumor Microenvironment

It has been shown that exosomes may be derived from different types of cancer cells. Such tumor-derived exosomes (TDEs) can play an important role in cancer progression and metastasis by decreasing cytotoxicity and promoting a pro-tumorigenic microenvironment [32]. Their immune suppressive cargo prevents the antitumor immune response, thus reducing the effectiveness of cancer treatment (Figure 2). However, TDEs act as a double-edged sword, as they may also present tumor antigens promoting anti-tumor immune responses [33].
This is of great therapeutic interest, as exosomes can be engineered to express receptors enhancing the specificity of their target cell binding, increasing their efficacy, and reducing deleterious off-target effects [34]. Moreover, attempts to anchor interleukins (IL), e.g., IL-12, to the exosome membrane with glycosylphosphatidylinositol anchor technology showed great results in terms of inducing T-cell proliferation (Figure 2). Higher IFN-γ release and higher cytotoxic effects were also observed in comparison to untreated exosomes [35]. A study by Zhang et al., revealed that renal cell carcinoma-derived exosomes could deliver lncRNAs (Figure 1), namely, lncARSR, inducing macrophage polarization by activating the STAT3 signaling pathway that causes changes in cytokine secretion and phagocytosis, thereby promoting tumor development [28]. Other studies found that renal TDEs also downregulate JAK3 expression and STAT5 phosphorylation at high doses in T lymphocytes with no effect on JAK2. Similar to TDEs, renal cancer TDE-anchored IL-12 (EXO/IL-12) also provides a much smaller reduction in p-STAT5 expression but has the same inhibitory effect on JAK3, but not on JAK2. The induction of T lymphocyte cytotoxicity enhanced by EXO/IL-12 partly depends on the upregulation of p-STAT5 expression. In conclusion, EXO/IL-12 may phosphorylate STAT5 via an alternative approach to the JAK2/STAT5 pathway when TDEs suppress the JAK3/STAT5 pathway in T lymphocytes [35,36]. LncRNAs in TDEs associated with breast cancer brain metastasis boost JAK2 activity that activates STAT3. This leads to macrophage activation and recruitment, promoting breast cancer brain metastasis as exosomes penetrate the blood-brain barrier [37].
Studies have shown that the development of cancer cachexia correlates with exosome concentration [38]. C26 cell-secreted exosomes can reduce the C2C12 myotube diameter in vitro and even reduce tibialis anterior muscle weight along with mouse grip strength in vivo [39]. The expression levels of STAT3 and STAT3 in C26 cells and the exosome release from C26 cells were decreased by siRNA [40]. In C26 cells, the overexpression of the STAT3 upregulates exosome biogenesis in sync with the level of exosomal markers. These results foreshadow that the STAT3 expression is linked with the exosome biogenesis in the C26 tumor cell line. Examination of the exosomes with electron microscopy showed that the TDEs in this case have a cup-shaped morphology with a diameter of about 50.7 nm [38,41]. In comparison to the control cell group, the culture medium of the C26 knockdown mock cells and the C26 overexpressed mock groups indicated a significant decrease in the diameter of the myotubes [39]. On the other hand, the culture medium of the C26-STAT3 knockdown cells only induced a mild subside in the diameter; in contrast, the culture medium of the C26-STAT3-overexpressed cells showed an enhanced decrease. In C26 cells, the knock-out STAT3 significantly reduces glycolysis [38,42]. Inhibiting the STAT3 phosphorylation henceforward reduces the transcriptional levels of HIF-1α, HK2, and serine hydroxymethyltransferase 2 along with pyruvate kinase M2 (PKM2) activation, which affects the synaptosome-associated protein 23 (SNAP23) activation, thus regulating exosome biogenesis [43]. IL-6 has been reported to play a pivotal role in the induction of muscle atrophy and lipolysis. It is also known that IL-6 binds to its receptors forms a complex that activates the STAT3; thus, the STAT3 is phosphorylated via the JAK, and this activation is preferred by the PKM2, which leads to the promotion of serine hydroxymethyltransferase 2 transcription in cancer [44,45]. The dimerized form of the PKM2 and the STAT3 enhances the glycolytic effect, which leads to enhanced exosome secretion. On the contrary, the phosphorylated PKM2 performs as a protein kinase, thus phosphorylating the SNAP23 and consequently promoting the SNARE complex formation, and hence activating the exosome release [39,43]. TDEs could play a pivotal role in the induction of cancer cachexia [38].
Exosomes may also transfer small 19–24 nucleotide-long, non-coding RNAs called microRNAs (miRNAs) that regulate gene expression [46,47,48,49,50,51,52]. Lin et al., found that exosome-derived miR-3184-5p expression levels were lowered in gastric cancer (GC) patients. miR-3184-5p has been shown to bind X-box binding protein 1 (XBP-1), a potential transcription activator of STAT3 that plays a critical role in cancer transformation and carcinogenesis [25]. Thus, downregulated miR-3184-5p may inhibit gastric cancer cell proliferation (Figure 2). However, silencing of XBP-1 was also shown to decrease the expression of p65, p-AKT, and p-STAT3, thereby inducing cell apoptosis (Table 1) [23,24,25]. TDEs advance pro-tumor inflammation via toll-like receptor (TLR) signaling (Figure 2). In lung cancer, the tumor cell-secreted exosomal miR-21 and miR-29a showed that they could connect to the TLR-8, inducing the activation of the NF-kB that leads to the production of pro-inflammatory cytokines, metastasis, and tumor proliferation (Figure 2) [2]. The study by De Veirman et al., found that multiple myeloma cell-derived exosomes may deliver miR-146a, which may affect not only the JAK/STAT pathway but also the Notch (via targeting Numb, a repressor of the Notch signaling pathway) and MAPK pathways (Table 1) [3]. Numerous studies suggest a crosstalk between JAK/STAT and Notch signaling, with each pathway capable of suppressing the other (Table 1) [3,20,22]. In addition to the above-mentioned role of tumor-derived exosomes, HPV infection generates cervical cancer cells that secrete CXCL10 and CXCR3 complexes binding to fibroblasts and activating the JAK/STAT pathway, which leads to the upregulation of the exosomal programmed death-ligand 1 (PD-L1), creating resistance to cell death (Figure 2) and forming an immune escape response pathway that leads to tumor emergence [53].
Figure 2. Biological functions of TDEs in tumor microenvironment. Membrane-associated Hsp72 from TDEs mediates STAT3-dependent immunosuppressive function of myeloid-derived suppressor cells (MDSC) [54]. miR-3184-5p expression was downregulated in gastric cancer (GC) patients, resulting in increased cell proliferation, migration, and invasion and reduced cell apoptosis [25]. CXCL10 produced by HPV-positive cervical squamous cell carcinoma (CSCC) stimulates exosomal PDL1 expression by fibroblasts via CXCR3 and JAK-STAT pathways, forming an immune escape response pathway that leads to tumor emergence [53]. TDEs from renal cancer cells suppressed the expression of JAK3 and p-STAT5 in T lymphocytes, reducing T-cell proliferation and cytotoxic effects [35]. Membrane-associated Hsp72 from TDEs mediates STAT3-dependent immunosuppressive functions [2]. Created with BioRender.com.
Figure 2. Biological functions of TDEs in tumor microenvironment. Membrane-associated Hsp72 from TDEs mediates STAT3-dependent immunosuppressive function of myeloid-derived suppressor cells (MDSC) [54]. miR-3184-5p expression was downregulated in gastric cancer (GC) patients, resulting in increased cell proliferation, migration, and invasion and reduced cell apoptosis [25]. CXCL10 produced by HPV-positive cervical squamous cell carcinoma (CSCC) stimulates exosomal PDL1 expression by fibroblasts via CXCR3 and JAK-STAT pathways, forming an immune escape response pathway that leads to tumor emergence [53]. TDEs from renal cancer cells suppressed the expression of JAK3 and p-STAT5 in T lymphocytes, reducing T-cell proliferation and cytotoxic effects [35]. Membrane-associated Hsp72 from TDEs mediates STAT3-dependent immunosuppressive functions [2]. Created with BioRender.com.
Pharmaceutics 15 01904 g002

3. Potential Therapeutic Implication of JAK/STAT Pathway

3.1. Signal Transducer and Activator of Transcription Factors (STATs)

STATs are regulating tumor proliferation, differentiation, and apoptosis [55]. STATs consist of 750–900 amino acids and have two phosphorylation sites: one of them is tyrosine 705, which is necessary for the STAT activation, and the other one is serine 727, which is phosphorylated after the tyrosine and does not lead to transactivation by itself. These are essential for STAT dimerization and transactivation. The activation of STAT begins with the binding of extracellular signals such as epidermal growth factors (EGF) and cytokines, which leads to the dimerization of the receptors [56,57,58].

3.1.1. STAT1

STAT1 can be activated by cytokines, IFNs, IL-2, IL-6, different growth factors, tumor necrosis factor (TNF), and angiotensin II. STAT1 has numerous functions like regulation of cell cycle-related genes or inhibition of cyclin expression. These lead to the inhibition of cell growth [59,60]. STAT1 may also regulate cell differentiation through phosphorylation and promote apoptosis by expressing multiple apoptosis proteins [61].
In osteosarcoma tissues, collagen type VI alpha 1 (COL6A1) is highly expressed [62,63]. It has been revealed in other cancer types as well, like prostate cancer, renal cell carcinoma, and cervical cancer, and these are associated with poor survival outcomes [64]. Cancer-associated fibroblasts activate fibroblasts that reside in the TME and produce a large number of chemokines, growth factors like TGF-β, cytokines like IL-1β, IL-6, IL-8, and collagen [65]. COL6A1 can be packed into exosomes and internalized by cancer cells, while its overexpression was shown to improve cell motility and metastasis in pancreatic cancer [66]. STAT1 significantly increases cell cycle arrest and apoptosis and decreases the capability of colony formation, cell migration, and cell invasion by suppressing epithelial-mesenchymal transition in osteosarcoma cells with COL6A1 overexpression [67].

3.1.2. STAT2

STAT2 is mostly activated by type I IFNs, especially FN-α and IFN-β [68]. STAT2 is different from other members of the family, as it is the only STAT that is not capable of forming homopolymers and does not directly bind to DNA [69]. Its functions are the antiviral effects via interferon-stimulated genes and the regulation of immune responses [68]. The transcription factor STAT3 is initiated by IFN-I, IL-6, and IL-10 family members, along with IL-21, IL-27, G-CSF, and leptin [70,71,72,73]. It is the main negative immune regulator of cell growth, differentiation, apoptosis, immune response, and tumor metastasis, and is also responsible for signal transduction processes [74].

3.1.3. STAT3

The JAK/STAT3 pathway has a key role in the growth and development of many human cancers. Elevated levels of IL-6 are frequently observed in malignancies and stimulate hyperactivation of JAK/STAT3 signaling (by induction of VEGF, cyclin D1, and MMPs), often associated with poor patient outcomes [75]. STAT3 activation influences the expression of the vascular endothelial growth factor (VEGF) and the hypoxia-inducible factor 1-alpha (HIF1-α), which leads to angiogenesis (Figure 2) [63]. Moreover, exosome surface-expressed Hsp72 in myeloid-derived suppressor cells (MDSCs) also showed the ability to trigger STAT3 in a TLR-2/MyD88-dependent manner leading to autocrine IL-6 production [40].
Chen et al., found that hBM-MSCs promote GC growth by regulating c-Myc. In addition to GC development being regulated by c-Myc, it also takes place in breast cancer cell proliferation as it suppresses apoptosis via upregulating the genes cyclin D1 and c-MYC with the help of the STAT3 pathway [76]. The higher level of expression of the cyclin D1 increases the loss of expression of MSH2, which leads to the appearance of Lynch syndrome. The increased expression of the c-Myc influences the expression of the exosomal human telomerase reverse transcriptase (hTERT). The exosomal hTERT-mRNA level is more elevated in Lynch syndrome mutation carriers (MSH2, MSH6), especially in metastatic colon cancer [77]. Exosomes, with the help of STAT pathways, could be a novel intermediate in the part of MSCs in GC promotion, and MSC-derived exosomes may be an unexplored therapeutic target not only for GC and ovarian cancer treatment but also for novel diagnostics of Lynch syndrome [46,76,78,79,80].
Ham et al., found that glycoprotein 130 (gp130) found in TDEs activates the IL-6/STAT3 pathway in macrophages, thereby helping to form the tumor microenvironment [31,81]. Gp130 is a signal transduction receptor for a cytokine family, which is involved in JAK/STAT signaling [31,72,82]. TDEs can transfer gp130 onto the bone marrow-derived macrophage cell membrane, where it can further accumulate. The inhibition of exosomal gp130 may reverse the effects of TDEs on macrophages [31]. TDEs can transfer gp130 onto the bone marrow-derived macrophages cell membrane where it can accumulate. For this inhibition, the implementation of EDTA is a viable approach [9]. TDEs pre-treated with SC144 lower the exosome-activated STAT3 level and additionally decrease the bone marrow-derived macrophages’ nuclear translocalization level. The incubation of the MDMs with pre-treated TDEs changes the IL-6 secretion phenotype [83]. Likewise, pre-treated TDEs mediate a decrease in gp130/STAT3 gene expression. In addition, in the bone marrow-derived macrophages, TDE-generated morphological and pro-survival changes were reverted by the pre-treated TDEs. All results confirm that exosomal gp130 plays a pivotal role in the STAT3 activation and changes in IL-6 secretion, as well as in the enhancement of survival and morphological changes in bone marrow-derived macrophags reacting to TDEs [31,83]. The study by Ying et al., found that ovarian cancer cells may release exosomes that carry miR-222-3p to macrophages. miR-222-3p overexpression may induce the polarization of the M2 phenotype in macrophages (Table 1) [27]. This type produces IL-10 at high levels, thereby promoting cancer cell proliferation. This regulation is initiated by the STAT3 pathway [84]. Also, inhibition via exosomes of the post-KRAS shows decreased levels of miR-210 expression, the downstream target of which is STAT3 in lung- and pancreatic-related tumors [29,30,85,86]. Inhibiting the STAT3 pathway might have therapeutic potential in ovarian cancer, as this pathway plays a pivotal role in the migration, invasiveness, and proliferation of tumor cells [27,87].
Recent studies have demonstrated that exosomes generated by senescent neutrophils, which are commonly present in therapy-treated tissues, can induce a resistant phenotype in recipient cancer cells. Exosomal piRNA-17560 secreted by senescent neutrophils through the STAT3-dependent pathway plays a crucial role in inducing chemoresistance and promoting epithelial-mesenchymal transition in breast cancer cells. These insights offer a potential approach to enhance therapeutic effectiveness by leveraging exosomes derived from senescent neutrophils [88].

3.1.4. STAT4

STAT4 is activated by type I IFN, IL-12, and IL-23 [89]. STAT4 phosphorylation is needed for the humoral immune response; it activates the germinal center response when viruses invade the body [90]. It was demonstrated that miR-141 can target the STAT4 gene expression to inhibit the proliferation, migration, and invasion of liver cancer cells [91]. On the other hand, the pro-oncogenic role of miR-141/200c in mature T-cell lymphoma cells via the altered expression of genes regulating cell survival and differentiation, including STAT4, has been shown. Moreover, an association between miR-141/200c-driven downregulation of STAT4 with an immature phenotype and shortened survival in primary T-prolymphocytic leukemia cases has been demonstrated [92]. STAT4 is also a target of miR-155, and inhibition of this oncogenic miRNA leads to the upregulation of STAT4 in MyLa cells [93]. In such cases, using engineered exosomes to deliver miRNA mimics or miRNA sponges may serve as a potential therapeutic strategy.

3.1.5. STAT5

STAT5 is mainly activated by the IL-2 cytokine family and IL-3 prolactin, but it can also be activated by the plate-derived growth factors EGF, EPO, GM-CSF, TPO, and GH. STAT5′s main biological function is the regulation of tumor immunity, development, cell growth, differentiation, and apoptosis [94,95,96,97,98]. High doses of renal cancer-derived exosomes inhibit STAT5 phosphorylation in T lymphocytes. TDEs are also known to suppress natural killer (NK) cells’ recruitment, migration, proliferation, and survival by reducing phosphorylation of STAT5 [99,100]. There are two major therapeutic approaches to counter such effects: the removal of circulating TDEs or blocking their formation and release [101,102] and abolishing the adverse effects of TDEs on NK cells using IL-15 [103,104]. Apart from restoring the expression of NK group 2D receptors, IL-15 may also restore STAT5 phosphorylation levels in NK cells [105]. Ye et al., found that in nasopharyngeal carcinoma, which is an Epstein–Barr virus-associated malignancy, the increased concentration of exosomes is in tune with a developed lymphoid node stage and poor prognosis [106]. TW03-derived exosomes inhibit T-cell proliferation along with the differentiation of the Th1 and Th17 [106,107]. Moreover, they advance Treg induction via NPC cells [108]. The above-mentioned outcomes are associated with a drop in ERK, STAT1, and STAT3 phosphorylation and the advancement of the STAT5 phosphorylation in the exo-stimulated T-cells. Also, TW03-derived exosomes enrich the proinflammatory cytokines such as IL-1β, IL-6, and IL-10 but reduce the IL-2, and IL-17 and IFNγ release from CD4+ and CD8+ T-cells. In nasopharyngeal carcinoma, there are overexpressed miRNAs in the exosomes, which were isolated from patient serum and cells, namely, hsa-miR-24-3p, hsa-miR-891a, hsa-miR-106a-5p, hsa-miR-20a-5p, and hsa-miR-1908 [106].

3.1.6. STAT6

IL-4 and viruses may activate STAT6. The functions of STAT6 are to promote B cell proliferation and maturation and to mediate the expression of MHC II and IgE [109]. A novel approach in immunotherapy is checkpoint inhibition. In metastatic cancer, this method induces a long-lasting tumor response [110,111]. The most promising checkpoint immunotherapy is reprogramming immunosuppressive tumor-associated macrophages (TAMs) to generate antitumor immunity. However, within the tumor microenvironment, myeloid cells are the primary antagonist mechanism to checkpoint immunotherapy [112]. TAMs are myeloid subgroups in the TME that show an immunosuppressive phenotype like the M2 [110,113,114]. Clinical data indicate a strong association between TAMs and poor prognosis. The M2 phenotype is regulated by STAT6, which has an impact on the TAMs. Upon interleukin stimulation, activated STAT6 dimerizes and alters the nucleus, which influences the transcription of M2 signature genes, in addition to suppressing the activation of M1 or inflammatory genes [115,116,117]. A novel approach is exosome-based TAMs reprogramming. Others have also investigated this method which selectively distributes STAT6 targeting antisense oligonucleotides (ASO) to TAMs [118]. exoASO-STAT6, in the wake of intravenous administration in the liver, exhibits biodistribution and STAT6-silencing activity with minimal distribution to other tissues. exoASO-STAT6 displays potent antitumor activity in numerous preclinical tumor models by inducing remodeling of the tumor microenvironment. exoASO-STAT6 applied via intravenous or intratumoral in syngenic tumor models appertaining to colorectal cancer and hepatocellular carcinoma proceeded to significant inhibition of tumor growth and complete tumor remission [110,119].
To investigate the effect of non-small cell lung cancer cells on the differentiation of the M2 macrophages, MRC-5 cells, and H1299 cell lines were co-incubated with PMA-stimulated THP-1 cells [120]. The results showed an intensified M2 macrophage number in the H1299 cell culture medium [121]. GW4869 inhibits M2 and macrophage marker secretion, suggesting that in H1299 cells the differentiation of the M2 macrophage happens through exosome secretion [122]. Also, they found that the exosomes from H1299 cells are more homogeneous and more spherical than typical exosomes, with a diameter of around 99 nm [123]. H1299-derived exosomes diminished the level of M1 markers and raised the expression in the M2 markers. On the other hand, the pretreatment with an exosome inhibitor showed no impact on M1 and M2 markers [123]. Other studies have revealed that macrophage-specific target gene transcription is activated by STAT6, which results in the differentiation of the M2 macrophage [124,125]. LINC00313 is linked to the expression of the STAT6. M1 markers like INOS, CD86, TNF-α, and IL-1β have increased expression levels if LINC00313 is knocked down, along with reduced M2 markers like TGFβ expression, which is caused by the overexpression of the STAT6. The expression of the CD206 along with the CD163-positive M2 cells also decreases if the LINC00313 is knocked down, whereas it increases when the STAT6 is overexpressed. These results indicate that the LINC00313 expression has an impact on the differentiation of the M2 macrophages via the STAT6 upregulation [120].

3.2. Janus Kinase Signaling Pathways

There are four members of the JAK family: JAK1, JAK2, JAK3, and tyrosine kinase 2. These are non-receptor tyrosine protein kinases. JAK1 may be phosphorylated by cytokine subunits, such as γc receptors, class II receptors, and gp130 subunits [126]. JAK1 may phosphorylate all the members of the STAT family [127]. It is expressed in all tissues and may induce hematopoiesis after activation via IL-3, IL-5, and IL-7 [128]. Mesenchymal stroma/stem-like cells (MSCs) are known for their antitumor activity. MSC-derived exosomes induce tumor cell growth, metastasis, and invasion. Glycoprotein A repetition predominant (GARP) is an oncogene in breast cancer. Deregulated GARP has been reported to promote aggressive tumor biology. siGARP-MSC exosomes inhibit the secretion of IL-6 and also inactivate the JAK1/STAT3 axis [129,130,131,132,133]. As such exosomes also reduced the proliferation and invasion of a mouse colon cancer cell line, a group of authors have proposed them as a potential novel anti-cancer tool worth investigating in the human context [133].

3.2.1. JAK1

The JAK/STAT3 pathway is one of the modulators of the MSC’s immunosuppression [134]. Phosgen is a highly poisonous gas and an intermediate product in different types of medicines, plastics, and dyes [135,136]. Exposure to phosgens could cause acute lung injury and cancer. To this day, there is no effective treatment to reverse the damage induced by phosgen [137]. Lung-derived exosomes upregulate the expression of the miR-28-5p, which promotes the MSCs function in phosgen induce acute lung injury. On the other hand, lung-derived exosomes reduce the manifestation of the miR-34c-3p, which activates the MSCs function via the JAK1/STAT3 pathway [138]. Suppressors of cytokine signaling (SOCS) are intracellular proteins that negatively regulate the JAK/STAT pathway as well as inflammation and immune responses [139]. Several studies have targeted the inhibition or activation of the JAK/STAT pathway’s different movements. Masoumi-Dehghi et al., found that the ovarian tumor cell-derived small extracellular vesicles upregulate the JAK-STAT pathway via down-regulating the SOCS5 [140,141]. Notably, the inhibitory effect of these extracellular vesicles on the SOCS5 expression level was partially rescued when transfected with anti-miR-141. It is indicated that JAK1 and STAT3 transcript expression levels were dramatically enhanced in TD-sEV-stimulated HUVECs; however, the stimulating impact of TD-sEVs on JAK1 and STAT3 expression levels was somewhat inhibited in the presence of anti-miR-141-3p. Also, the reduced SOCS5 expression resulting from small extracellular vesicles transfer of tumor-secreted miR-141-3p upregulates the expression of VEGFR-2 by stimulating the JAK/STAT signaling pathway in endothelial cells [142]. The ovarian tumor cell-derived small extracellular vesicles upregulate the VEGFR-2, thereby promoting endothelial cell migration and angiogenesis in vitro [139,143].

3.2.2. JAK2

JAK2 is also phosphorylated by the cytokine subunit, gp130 receptor family, and the class II cytokine receptor family. It is a signal transductor of receptors of IL-3, growth hormone, and prolactin [144]. Zhang et al., found that renal cancer-derived exosomes downregulated the expression of JAK3 but there was no effect on the expression of JAK2 in T-lymphocytes (Figure 2) [35]. Exosomes and tumor-associated macrophages have complex crosstalk in the glioma microenvironment [145]. Ming et al., studied the glioma progression and the macrophage polarization correlation with the effect of the TDE [146]. Firstly, they looked for functional miRNAs that can only be found in glioma by sequencing the cerebrospinal fluid and glioma tissues. miR-3591-3p met the criteria perfectly; it can also significantly influence the M2 macrophage polarization and the secretion of the IL10 and TGFB1, which leads to glioma metastasis, and its overexpression induces apoptosis [147]. It can also activate different pathways like the JAK2 pathway and STAT3 pathway. Lentivirally transduced macrophages via miR-3591-3p significantly stimulate glioma progression. This study revealed that tumor-suppressive miR-3591-3p is secreted by exosomes in glioma cells and targets tumor-associated macrophages to generate an immunosuppressive microenvironment [146,148,149]. The next step was to examine the expression level of the miR-3591-3p in exosomes by examining it in different glioma and normal astrocyte cell lines. Increased expression level of exosomal miR-3591-3p was found in the studied glioma cell lines. Investigating the influence of glioma cells on macrophages in the tumor microenvironment, they co-cultured glioma cells THP-1-Mφ yielded with PMA in vitro. In these groups, they gauged the M1 and M2 macrophage-associated phenotypic marker levels via qRT-PCR. The level of M2 markers was significantly increased in the glioma co-cultured groups, on the other hand, the exosome release inhibitor-treated group’s M1 markers decreased [146].
Chim et al., studied SOCS1 and SHP1 hypermethylation in multiple myeloma. SOCS1 may bind to the JAK2 kinase SH2 domain, inhibiting its activity, and is also a negative regulator of IL-6; therefore, it may function as a tumor suppressor in multiple myeloma. SHP1 is also a negative regulator of the JAK/STAT signaling pathway. The studied genes were shown to be inactive due to methylation; therefore, demethylating agents may be potential therapeutic agents [21].

3.2.3. JAK3

JAK3 expression is only limited to hematopoietic progenitor cells and is mainly required for γc receptor signal transduction. JAK3 deregulation may lead to combined autosomal recessive immunodeficiency [150,151]. Differential microarray analysis showed that in prostate cancer tissue, the miR-187 expression decreased, which is correlated with poor prognosis in prostate cancer patients. On the other hand, prostate cancer cells’ malignant phenotype is inhibited by miR-187 overexpression [26]. The JAK3/STAT3-slug axis, which is reduced by human bone marrow-derived mesenchymal stem cell-derived exosomes miR-187, maintains prostate cancer metastasis and cell growth (Table 1) [26]. In multiple myeloma cells, metastasis can advance by CD276 via activation of the JAK/STAT pathway [152]. Overexpressed CD276 has been reported in numerous malignant tumors, which were reported to advance the metastasis of cancer cells. Overexpressed CD278 correlates with poor clinical prognosis [26,152].

4. Conclusions

This review has shown the promising roles of exosomes in diagnostic and therapeutic applications. These vesicles provide a potential tool for assessing tumor progression and malignant state; however, they may also be engineered to complement antitumor therapy. Several strategies to modify the tumor microenvironment have been suggested, including the use of TDE membrane-anchored interleukins and TDE-encapsulated miRNAs affecting the JAK/STAT signaling pathway. Exosome processing and analysis were not addressed in this review, but since the workflow has not been standardized yet, the methodology varies between studies, which may result in inconsistent findings. Nevertheless, exosomes and TDE modifications show clear promise for personalized diagnostics and therapy.

Author Contributions

Conceptualization, B.S. and G.G.; writing—original draft preparation, G.G., N.N., O.P., J.S. and B.S.; writing—review and editing, G.G., N.N., O.P., J.S., G.B., L.D., B.N., I.B. and B.S.; visualization, G.G., O.P., J.S. and B.S.; supervision, B.N., I.B. and B.S.; funding acquisition, T.S. All authors have read and agreed to the published version of the manuscript.

Funding

Funding was provided by the Operational Programme Integrated Infrastructure for the project ITMS: 313011V578 (PreveLynch), co-financed by the European Regional Development Fund.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lin, B.; Lei, Y.; Wang, J.; Zhu, L.; Wu, Y.; Zhang, H.; Wu, L.; Zhang, P.; Yang, C. Microfluidic-Based Exosome Analysis for Liquid Biopsy. Small Methods 2021, 5, e2001131. [Google Scholar] [CrossRef]
  2. Fabbri, M.; Paone, A.; Calore, F.; Galli, R.; Gaudio, E.; Santhanam, R.; Lovat, F.; Fadda, P.; Mao, C.; Nuovo, G.J.; et al. MicroRNAs bind to Toll-like receptors to induce prometastatic inflammatory response. Proc. Natl. Acad. Sci. USA 2012, 109, E2110-6. [Google Scholar] [CrossRef] [PubMed]
  3. De Veirman, K.; Wang, J.; Xu, S.; Leleu, X.; Himpe, E.; Maes, K.; De Bruyne, E.; Van Valckenborgh, E.; Vanderkerken, K.; Menu, E.; et al. Induction of miR-146a by multiple myeloma cells in mesenchymal stromal cells stimulates their pro-tumoral activity. Cancer Lett. 2016, 377, 17–24. [Google Scholar] [CrossRef] [PubMed]
  4. Xie, S.; Zhang, Q.; Jiang, L. Current Knowledge on Exosome Biogenesis, Cargo-Sorting Mechanism and Therapeutic Implications. Membranes 2022, 12, 498. [Google Scholar] [CrossRef] [PubMed]
  5. Schmidt, O.; Teis, D. The ESCRT machinery. Curr. Biol. 2012, 22, R116–R120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Hessvik, N.P.; Llorente, A. Current knowledge on exosome biogenesis and release. Cell. Mol. Life Sci. 2018, 75, 193–208. [Google Scholar] [CrossRef] [Green Version]
  7. Budnik, V.; Ruiz-Cañada, C.; Wendler, F. Extracellular vesicles round off communication in the nervous system. Nat. Rev. Neurosci. 2016, 17, 160–172. [Google Scholar] [CrossRef] [Green Version]
  8. Gerlach, J.Q.; Griffin, M.D. Getting to know the extracellular vesicle glycome. Mol. Biosyst. 2016, 12, 1071–1081. [Google Scholar] [CrossRef]
  9. Barrès, C.; Blanc, L.; Bette-Bobillo, P.; André, S.; Mamoun, R.; Gabius, H.-J.; Vidal, M. Galectin-5 is bound onto the surface of rat reticulocyte exosomes and modulates vesicle uptake by macrophages. Blood 2010, 115, 696–705. [Google Scholar] [CrossRef] [Green Version]
  10. Domenis, R.; Cesselli, D.; Toffoletto, B.; Bourkoula, E.; Caponnetto, F.; Manini, I.; Beltrami, A.P.; Ius, T.; Skrap, M.; Di Loreto, C.; et al. Systemic T Cells Immunosuppression of Glioma Stem Cell-Derived Exosomes Is Mediated by Monocytic Myeloid-Derived Suppressor Cells. PLoS ONE 2017, 12, e0169932. [Google Scholar] [CrossRef] [Green Version]
  11. Mathieu, M.; Névo, N.; Jouve, M.; Valenzuela, J.I.; Maurin, M.; Verweij, F.J.; Palmulli, R.; Lankar, D.; Dingli, F.; Loew, D.; et al. Specificities of exosome versus small ectosome secretion revealed by live intracellular tracking of CD63 and CD9. Nat. Commun. 2021, 12, 4839. [Google Scholar] [CrossRef]
  12. Potolicchio, I.; Carven, G.J.; Xu, X.; Stipp, C.; Riese, R.J.; Stern, L.J.; Santambrogio, L. Proteomic Analysis of Microglia-Derived Exosomes: Metabolic Role of the Aminopeptidase CD13 in Neuropeptide Catabolism. J. Immunol. 2005, 175, 2237–2243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Kim, K.M.; Abdelmohsen, K.; Mustapic, M.; Kapogiannis, D.; Gorospe, M. RNA in extracellular vesicles. Wiley Interdiscip. Rev. RNA 2017, 8, e1413. [Google Scholar] [CrossRef]
  14. Mathivanan, S.; Ji, H.; Simpson, R.J. Exosomes: Extracellular organelles important in intercellular communication. J. Proteom. 2010, 73, 1907–1920. [Google Scholar] [CrossRef] [PubMed]
  15. Dorayappan, K.D.P.; Wanner, R.; Wallbillich, J.J.; Saini, U.; Zingarelli, R.; Suarez, A.A.; Cohn, D.E.; Selvendiran, K. Hypoxia-induced exosomes contribute to a more aggressive and chemoresistant ovarian cancer phenotype: A novel mechanism linking STAT3/Rab proteins. Oncogene 2018, 37, 3806–3821. [Google Scholar] [CrossRef]
  16. Pauletto, E.; Eickhoff, N.; Padrão, N.A.; Blattner, C.; Zwart, W. TRIMming Down Hormone-Driven Cancers: The Biological Impact of TRIM Proteins on Tumor Development, Progression and Prognostication. Cells 2021, 10, 1517. [Google Scholar] [CrossRef]
  17. Yu, X.-Z.; Yuan, J.-L.; Ye, H.; Yi, K.; Qie, M.-R.; Hou, M.-M. TRIM44 facilitates ovarian cancer proliferation, migration, and invasion by inhibiting FRK. Neoplasma 2021, 68, 751–759. [Google Scholar] [CrossRef] [PubMed]
  18. Ye, H.; Wang, R.; Yu, X.; Wu, Y.; Yang, B.; Ao, M.; Xi, M.; Hou, M. Exosomal circNFIX promotes angiogenesis in ovarian cancer via miR -518a-3p/TRIM44 axis. Kaohsiung J. Med. Sci. 2022, 39, 26–39. [Google Scholar] [CrossRef]
  19. Cao, H.; Gao, R.; Chen, L.; Feng, Y. TRIM66 promotes malignant progression of prostate carcinoma through the JAK/STAT pathway. FEBS Open Bio 2020, 10, 515–524. [Google Scholar] [CrossRef] [Green Version]
  20. Cowan, A.J.; Green, D.J.; Kwok, M.; Lee, S.; Coffey, D.G.; Holmberg, L.A.; Tuazon, S.; Gopal, A.K.; Libby, E.N. Diagnosis and Management of Multiple Myeloma. JAMA 2022, 327, 464–477. [Google Scholar] [CrossRef]
  21. Chim, C.-S.; Fung, T.-K.; Cheung, W.-C.; Liang, R.; Kwong, Y.-L. SOCS1 and SHP1 hypermethylation in multiple myeloma: Implications for epigenetic activation of the Jak/STAT pathway. Blood 2004, 103, 4630–4635. [Google Scholar] [CrossRef]
  22. Hu, X.; Li, J.; Fu, M.; Zhao, X.; Wang, W. The JAK/STAT signaling pathway: From bench to clinic. Signal Transduct. Target. Ther. 2021, 6, 402. [Google Scholar] [CrossRef]
  23. Wagner, A.D.; Syn, N.L.; Moehler, M.; Grothe, W.; Yong, W.P.; Tai, B.-C.; Ho, J.; Unverzagt, S. Chemotherapy for advanced gastric cancer. Cochrane Database Syst. Rev. 2017, 2017, CD004064. [Google Scholar] [CrossRef] [Green Version]
  24. Smyth, E.C.; Nilsson, M.; Grabsch, H.I.; van Grieken, N.C.; Lordick, F. Gastric cancer. Lancet 2020, 396, 635–648. [Google Scholar] [CrossRef]
  25. Lin, S.; Que, Y.; Que, C.; Li, F.; Deng, M.; Xu, D. Exosome miR-3184-5p inhibits gastric cancer growth by targeting XBP1 to regulate the AKT, STAT3, and IRE1 signalling pathways. Asia-Pacific J. Clin. Oncol. 2022, 19, e27–e38. [Google Scholar] [CrossRef]
  26. Li, C.; Sun, Z.; Song, Y.; Zhang, Y. Suppressive function of bone marrow-derived mesenchymal stem cell-derived exosomal microRNA-187 in prostate cancer. Cancer Biol. Ther. 2022, 23, 1–14. [Google Scholar] [CrossRef] [PubMed]
  27. Ying, X.; Wu, Q.; Wu, X.; Zhu, Q.; Wang, X.; Jiang, L.; Chen, X.; Wang, X. Epithelial ovarian cancer-secreted exosomal miR-222-3p induces polarization of tumor-associated macrophages. Oncotarget 2016, 7, 43076–43087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Zhang, W.; Zheng, X.; Yu, Y.; Zheng, L.; Lan, J.; Wu, Y.; Liu, H.; Zhao, A.; Huang, H.; Chen, W. Renal cell carcinoma-derived exosomes deliver lncARSR to induce macrophage polarization and promote tumor progression via STAT3 pathway. Int. J. Biol. Sci. 2022, 18, 3209–3222. [Google Scholar] [CrossRef]
  29. Kalvala, A.; Wallet, P.; Yang, L.; Wang, C.; Li, H.; Nam, A.; Nathan, A.; Mambetsariev, I.; Poroyko, V.; Gao, H.; et al. Phenotypic Switching of Naïve T Cells to Immune-Suppressive Treg-Like Cells by Mutant KRAS. J. Clin. Med. 2019, 8, 1726. [Google Scholar] [CrossRef] [Green Version]
  30. Dai, E.; Han, L.; Liu, J.; Xie, Y.; Kroemer, G.; Klionsky, D.J.; Zeh, H.J.; Kang, R.; Wang, J.; Tang, D. Autophagy-dependent ferroptosis drives tumor-associated macrophage polarization via release and uptake of oncogenic KRAS protein. Autophagy 2020, 16, 2069–2083. [Google Scholar] [CrossRef] [PubMed]
  31. Ham, S.; Lima, L.G.; Chai, E.P.Z.; Muller, A.; Lobb, R.J.; Krumeich, S.; Wen, S.W.; Wiegmans, A.P.; Möller, A. Breast Cancer-Derived Exosomes Alter Macrophage Polarization via gp130/STAT3 Signaling. Front. Immunol. 2018, 9, 871. [Google Scholar] [CrossRef] [Green Version]
  32. Whiteside, T.L. Tumor-Derived Exosomes and Their Role in Cancer Progression. Adv. Clin. Chem. 2016, 74, 103–141. [Google Scholar] [CrossRef] [Green Version]
  33. Hussain, W.A.; Jahangir, S.; Ghosh, B.; Yesmin, F.; Anis, A.; Satil, S.N.; Anwar, F.; Rashid, M.H. Exosomes for Regulation of Immune Responses and Immunotherapy. J. Nanotheranostics 2022, 3, 55–85. [Google Scholar] [CrossRef]
  34. Li, Q.; Wang, H.; Peng, H.; Huyan, T.; Cacalano, N.A. Exosomes: Versatile nano mediators of immune regulation. Cancers 2019, 11, 1557. [Google Scholar] [CrossRef] [Green Version]
  35. Zhang, Y.; Wu, X.-H.; Luo, C.-L.; Zhang, J.-M.; He, B.-C.; Chen, G. Interleukin-12-anchored exosomes increase cytotoxicity of T lymphocytes by reversing the JAK/STAT pathway impaired by tumor-derived exosomes. Int. J. Mol. Med. 2010, 25, 695–700. [Google Scholar] [CrossRef]
  36. Gao, Y.; Xu, H.; Li, N.; Wang, H.; Ma, L.; Chen, S.; Liu, J.; Zheng, Y.; Zhang, Y. Renal cancer-derived exosomes induce tumor immune tolerance by MDSCs-mediated antigen-specific immunosuppression. Cell Commun. Signal. 2020, 18, 106. [Google Scholar] [CrossRef] [PubMed]
  37. Ni, C.; Fang, Q.-Q.; Chen, W.-Z.; Jiang, J.-X.; Jiang, Z.; Ye, J.; Zhang, T.; Yang, L.; Meng, F.-B.; Xia, W.-J.; et al. Breast cancer-derived exosomes transmit lncRNA SNHG16 to induce CD73+γδ1 Treg cells. Signal Transduct. Target. Ther. 2020, 5, 41. [Google Scholar] [CrossRef] [PubMed]
  38. Fan, M.; Sun, W.; Gu, X.; Lu, S.; Shen, Q.; Liu, X.; Zhang, X. The critical role of STAT3 in biogenesis of tumor-derived exosomes with potency of inducing cancer cachexia in vitro and in vivo. Oncogene 2022, 41, 1050–1062. [Google Scholar] [CrossRef] [PubMed]
  39. Miao, C.; Zhang, W.; Feng, L.; Gu, X.; Shen, Q.; Lu, S.; Fan, M.; Li, Y.; Guo, X.; Ma, Y.; et al. Cancer-derived exosome miRNAs induce skeletal muscle wasting by Bcl-2-mediated apoptosis in colon cancer cachexia. Mol. Ther.-Nucleic Acids 2021, 24, 923–938. [Google Scholar] [CrossRef]
  40. Liu, Q.; Dai, G.; Wu, Y.; Zhang, M.; Yang, M.; Wang, X.; Song, M.; Li, X.; Xia, R.; Wu, Z. iRGD-modified exosomes-delivered BCL6 siRNA inhibit the progression of diffuse large B-cell lymphoma. Front. Oncol. 2022, 12, 822805. [Google Scholar] [CrossRef]
  41. Jung, M.K.; Mun, J.Y. Sample Preparation and Imaging of Exosomes by Transmission Electron Microscopy. J. Vis. Exp. 2018, 131, e56482. [Google Scholar] [CrossRef]
  42. Li, M.; Jin, R.; Wang, W.; Zhang, T.; Sang, J.; Li, N.; Han, Q.; Zhao, W.; Li, C.; Liu, Z. STAT3 regulates glycolysis via targeting hexokinase 2 in hepatocellular carcinoma cells. Oncotarget 2017, 8, 24777–24784. [Google Scholar] [CrossRef] [Green Version]
  43. Wei, Y.; Wang, D.; Jin, F.; Bian, Z.; Li, L.; Liang, H.; Li, M.; Shi, L.; Pan, C.; Zhu, D.; et al. Pyruvate kinase type M2 promotes tumour cell exosome release via phosphorylating synaptosome-associated protein 23. Nat. Commun. 2017, 8, 14041. [Google Scholar] [CrossRef] [Green Version]
  44. Bi, Y.-H.; Han, W.-Q.; Li, R.-F.; Wang, Y.-J.; Du, Z.-S.; Wang, X.-J.; Jiang, Y. Signal transducer and activator of transcription 3 promotes the Warburg effect possibly by inducing pyruvate kinase M2 phosphorylation in liver precancerous lesions. World J. Gastroenterol. 2019, 25, 1936–1949. [Google Scholar] [CrossRef] [PubMed]
  45. Gao, S.; Chen, M.; Wei, W.; Zhang, X.; Zhang, M.; Yao, Y.; Lv, Y.; Ling, T.; Wang, L.; Zou, X. Crosstalk of mTOR/PKM2 and STAT3/c-Myc signaling pathways regulate the energy metabolism and acidic microenvironment of gastric cancer. J. Cell. Biochem. 2018, 120, 1193–1202. [Google Scholar] [CrossRef]
  46. Ambros, V. microRNAs: Tiny Regulators with Great Potential. Cell 2001, 107, 823–826. [Google Scholar] [CrossRef] [Green Version]
  47. Bartel, D.P. MicroRNAs: Target Recognition and Regulatory Functions. Cell 2009, 136, 215–233. [Google Scholar] [CrossRef] [Green Version]
  48. Croce, C.M. Causes and consequences of microRNA dysregulation in cancer. Nat. Rev. Genet. 2009, 10, 704–714. [Google Scholar] [CrossRef] [PubMed]
  49. Fabbri, M.; Croce, C.M. Role of microRNAs in lymphoid biology and disease. Curr. Opin. Hematol. 2011, 18, 266–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Lawrie, C.H.; Gal, S.; Dunlop, H.M.; Pushkaran, B.; Liggins, A.P.; Pulford, K.; Banham, A.H.; Pezzella, F.; Boultwood, J.; Wainscoat, J.S.; et al. Detection of elevated levels of tumour-associated microRNAs in serum of patients with diffuse large B-cell lymphoma. Br. J. Haematol. 2008, 141, 672–675. [Google Scholar] [CrossRef] [PubMed]
  51. Gibbings, D.J.; Ciaudo, C.; Erhardt, M.; Voinnet, O. Multivesicular bodies associate with components of miRNA effector complexes and modulate miRNA activity. Nat. Cell Biol. 2009, 11, 1143–1149. [Google Scholar] [CrossRef] [PubMed]
  52. Mitchell, P.S.; Parkin, R.K.; Kroh, E.M.; Fritz, B.R.; Wyman, S.K.; Pogosova-Agadjanyan, E.L.; Peterson, A.; Noteboom, J.; O’Briant, K.C.; Allen, A.; et al. Circulating microRNAs as stable blood-based markers for cancer detection. Proc. Natl. Acad. Sci. USA 2008, 105, 10513–10518. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, X.; He, H.; Xiao, Y.; Hasim, A.; Yuan, J.; Ye, M.; Li, X.; Hao, Y.; Guo, X. CXCL10 Produced by HPV-Positive Cervical Cancer Cells Stimulates Exosomal PDL1 Expression by Fibroblasts via CXCR3 and JAK-STAT Pathways. Front. Oncol. 2021, 11, 629350. [Google Scholar] [CrossRef] [PubMed]
  54. Chalmin, F.; Ladoire, S.; Mignot, G.; Vincent, J.; Bruchard, M.; Remy-Martin, J.-P.; Boireau, W.; Rouleau, A.; Simon, B.; Lanneau, D.; et al. Membrane-associated Hsp72 from tumor-derived exosomes mediates STAT3-dependent immunosuppressive function of mouse and human myeloid-derived suppressor cells. J. Clin. Investig. 2010, 120, 457–471. [Google Scholar] [CrossRef]
  55. Gao, P.; Niu, N.; Wei, T.; Tozawa, H.; Chen, X.; Zhang, C.; Zhang, J.; Wada, Y.; Kapron, C.M.; Liu, J. The roles of signal transducer and activator of transcription factor 3 in tumor angiogenesis. Oncotarget 2017, 8, 69139–69161. [Google Scholar] [CrossRef] [Green Version]
  56. Lim, C.P.; Cao, X. Structure, function, and regulation of STAT proteins. Mol. Biosyst. 2006, 2, 536–550. [Google Scholar] [CrossRef]
  57. Zhu, X.; Zhou, W. The Emerging Regulation of VEGFR-2 in Triple-Negative Breast Cancer. Front. Endocrinol. 2015, 6, 159. [Google Scholar] [CrossRef] [Green Version]
  58. Banerjee, K.; Resat, H. Constitutive activation of STAT3 in breast cancer cells: A review. Int. J. Cancer 2015, 138, 2570–2578. [Google Scholar] [CrossRef] [Green Version]
  59. Schlee, M.; Hölzel, M.; Bernard, S.; Mailhammer, R.; Schuhmacher, M.; Reschke, J.; Eick, D.; Marinkovic, D.; Wirth, T.; Rosenwald, A.; et al. c-MYC activation impairs the NF-κB and the interferon response: Implications for the pathogenesis of Burkitt’s lymphoma. Int. J. Cancer 2007, 120, 1387–1395. [Google Scholar] [CrossRef]
  60. Dimberg, A.; Nilsson, K.; Oberg, F. Phosphorylation-deficient Stat1 inhibits retinoic acid–induced differentiation and cell cycle arrest in U-937 monoblasts. Blood 2000, 96, 2870–2878. [Google Scholar] [CrossRef]
  61. Lee, C.-K.; Smith, E.; Gimeno, R.; Gertner, R.; Levy, D.E. STAT1 Affects Lymphocyte Survival and Proliferation Partially Independent of Its Role Downstream of IFN-γ. J. Immunol. 2000, 164, 1286–1292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Zhang, Y.; Liu, Z.; Yang, X.; Lu, W.; Chen, Y.; Lin, Y.; Wang, J.; Lin, S.; Yun, J.-P. H3K27 acetylation activated-COL6A1 promotes osteosarcoma lung metastasis by repressing STAT1 and activating pulmonary cancer-associated fibroblasts. Theranostics 2021, 11, 1473–1492. [Google Scholar] [CrossRef] [PubMed]
  63. Kuo, H.-J.; Maslen, C.L.; Keene, D.R.; Glanville, R.W. Type VI Collagen Anchors Endothelial Basement Membranes by Interacting with Type IV Collagen. J. Biol. Chem. 1997, 272, 26522–26529. [Google Scholar] [CrossRef] [Green Version]
  64. Park, J.; Scherer, P.E. Adipocyte-derived endotrophin promotes malignant tumor progression. J. Clin. Investig. 2012, 122, 4243–4256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Gaggioli, C.; Hooper, S.; Hidalgo-Carcedo, C.; Grosse, R.; Marshall, J.F.; Harrington, K.; Sahai, E. Fibroblast-led collective invasion of carcinoma cells with differing roles for RhoGTPases in leading and following cells. Nat. Cell Biol. 2007, 9, 1392–1400. [Google Scholar] [CrossRef] [PubMed]
  66. Owusu-Ansah, K.G.; Song, G.; Chen, R.; Edoo, M.I.A.; Li, J.; Chen, B.; Wu, J.; Zhou, L.; Xie, H.; Jiang, D.; et al. COL6A1 promotes metastasis and predicts poor prognosis in patients with pancreatic cancer. Int. J. Oncol. 2019, 55, 391–404. [Google Scholar] [CrossRef] [Green Version]
  67. Sherman-Baust, C.A.; Weeraratna, A.T.; Rangel, L.B.; Pizer, E.S.; Cho, K.R.; Schwartz, D.R.; Shock, T.; Morin, P.J. Remodeling of the extracellular matrix through overexpression of collagen VI contributes to cisplatin resistance in ovarian cancer cells. Cancer Cell 2003, 3, 377–386. [Google Scholar] [CrossRef] [Green Version]
  68. Wang, Y.; Song, Q.; Huang, W.; Lin, Y.; Wang, X.; Wang, C.; Willard, B.; Zhao, C.; Nan, J.; Holvey-Bates, E.; et al. A virus-induced conformational switch of STAT1-STAT2 dimers boosts antiviral defenses. Cell Res. 2020, 31, 206–218. [Google Scholar] [CrossRef]
  69. Schindler, C.; Plumlee, C. Inteferons pen the JAK–STAT pathway. Semin. Cell Dev. Biol. 2008, 19, 311–318. [Google Scholar] [CrossRef] [Green Version]
  70. Zhong, Z.; Wen, Z.; Darnell, J.E., Jr. Stat3: A STAT Family Member Activated by Tyrosine Phosphorylation in Response to Epidermal Growth Factor and Interleukin-6. Science 1994, 264, 95–98. [Google Scholar] [CrossRef]
  71. Darnell, J.E., Jr.; Kerr, I.M.; Stark, G.R. Jak-STAT pathways and transcriptional activation in response to IFNs and other extracellular signaling proteins. Science 1994, 264, 1415–1421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Heinrich, P.C.; Behrmann, I.; Müller-Newen, G.; Schaper, F.; Graeve, L. Interleukin-6-type cytokine signalling through the gp130/Jak/STAT pathway. Biochem. J. 1998, 334, 297–314. [Google Scholar] [CrossRef] [Green Version]
  73. Ruff-Jamison, S.; Zhong, Z.; Wen, Z.; Chen, K.; Darnell, J.; Cohen, S. Epidermal growth factor and lipopolysaccharide activate Stat3 transcription factor in mouse liver. J. Biol. Chem. 1994, 269, 21933–21935. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, M.; Zhou, L.; Xu, Y.; Yang, M.; Xu, Y.; Komaniecki, G.P.; Kosciuk, T.; Chen, X.; Lu, X.; Zou, X.; et al. A STAT3 palmitoylation cycle promotes TH17 differentiation and colitis. Nature 2020, 586, 434–439. [Google Scholar] [CrossRef] [PubMed]
  75. Johnson, D.E.; O’Keefe, R.A.; Grandis, J.R. Targeting the IL-6/JAK/STAT3 signalling axis in cancer. Nat. Rev. Clin. Oncol. 2018, 15, 234–248. [Google Scholar] [CrossRef]
  76. Chen, D.; Liu, S.; Ma, H.; Liang, X.; Ma, H.; Yan, X.; Yang, B.; Wei, J.; Liu, X. Paracrine factors from adipose-mesenchymal stem cells enhance metastatic capacity through Wnt signaling pathway in a colon cancer cell co-culture model. Cancer Cell Int. 2015, 15, 42. [Google Scholar] [CrossRef] [Green Version]
  77. Buglyó, G.; Styk, J.; Pös, O.; Csók, Á.; Repiska, V.; Soltész, B.; Szemes, T.; Nagy, B. Liquid Biopsy as a Source of Nucleic Acid Biomarkers in the Diagnosis and Management of Lynch Syndrome. Int. J. Mol. Sci. 2022, 23, 4284. [Google Scholar] [CrossRef]
  78. Yang, Y.; Yang, J.-J.; Tao, H.; Jing-Jing, Y. MicroRNA-21 controls hTERT via PTEN in human colorectal cancer cell proliferation. J. Physiol. Biochem. 2015, 71, 59–68. [Google Scholar] [CrossRef]
  79. Marcolino, T.F.; Pimenta, C.A.M.; Neto, R.A.; Castelo, P.M.; Silva, M.S.; Forones, N.M.; Oshima, C.T.F. p53, Cyclin-D1, β-catenin, APC and c-myc in Tumor Tissue from Colorectal and Gastric Cancer Patients with Suspected Lynch Syndrome by the Bethesda Criteria. Asian Pac. J. Cancer Prev. 2020, 21, 343–348. [Google Scholar] [CrossRef]
  80. Wang, N.; Pei, B.; Yuan, X.; Yi, C.; Ocansey, D.K.W.; Qian, H.; Mao, F. Emerging roles of mesenchymal stem cell-derived exosomes in gastrointestinal cancers. Front. Bioeng. Biotechnol. 2022, 10, 1019459. [Google Scholar] [CrossRef]
  81. Sevko, A.; Umansky, V. Myeloid-Derived Suppressor Cells Interact with Tumors in Terms of Myelopoiesis, Tumorigenesis and Immunosuppression: Thick as Thieves. J. Cancer 2013, 4, 3–11. [Google Scholar] [CrossRef]
  82. Chow, D.-C.; He, X.-L.; Snow, A.L.; Rose-John, S.; Garcia, K.C. Structure of an Extracellular gp130 Cytokine Receptor Signaling Complex. Science 2001, 291, 2150–2155. [Google Scholar] [CrossRef]
  83. Bromberg, J.; Wang, T.C. Inflammation and Cancer: IL-6 and STAT3 Complete the Link. Cancer Cell 2009, 15, 79–80. [Google Scholar] [CrossRef] [Green Version]
  84. Deng, G.; Li, K.; Chen, S.; Chen, P.; Zheng, H.; Yu, B.; Zhang, K. Interleukin-10 promotes proliferation and migration, and inhibits tendon differentiation via the JAK/Stat3 pathway in tendon-derived stem cells in vitro. Mol. Med. Rep. 2018, 18, 5044–5052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Chen, Q.; Xie, X. Association of Exosomal miR-210 with Signaling Pathways Implicated in Lung Cancer. Genes 2021, 12, 1248. [Google Scholar] [CrossRef]
  86. Wang, Z.-Y.; Wang, R.-X.; Ding, X.-Q.; Zhang, X.; Pan, X.-R.; Tong, J.-H. A Protocol for Cancer-Related Mutation Detection on Exosomal DNA in Clinical Application. Front. Oncol. 2020, 10, 558106. [Google Scholar] [CrossRef]
  87. Liang, R.; Chen, X.; Chen, L.; Wan, F.; Chen, K.; Sun, Y.; Zhu, X. STAT3 signaling in ovarian cancer: A potential therapeutic target. J. Cancer 2020, 11, 837–848. [Google Scholar] [CrossRef] [Green Version]
  88. Ou, B.; Liu, Y.; Gao, Z.; Xu, J.; Yan, Y.; Li, Y.; Zhang, J. Senescent neutrophils-derived exosomal piRNA-17560 promotes chemoresistance and EMT of breast cancer via FTO-mediated m6A demethylation. Cell Death Dis. 2022, 13, 905. [Google Scholar] [CrossRef]
  89. Miyagi, T.; Gil, M.P.; Wang, X.; Louten, J.; Chu, W.-M.; Biron, C.A. High basal STAT4 balanced by STAT1 induction to control type 1 interferon effects in natural killer cells. J. Exp. Med. 2007, 204, 2383–2396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Weinstein, J.S.; Laidlaw, B.J.; Lu, Y.; Wang, J.K.; Schulz, V.P.; Li, N.; Herman, E.I.; Kaech, S.M.; Gallagher, P.G.; Craft, J. Correction: STAT4 and T-bet control follicular helper T cell development in viral infections. J. Exp. Med. 2018, 215, 999. [Google Scholar] [CrossRef] [PubMed]
  91. Ma, L.; Shao, H.; Chen, H.; Deng, Q. The Mechanism of miR-141 Regulating the Proliferation and Metastasis of Liver Cancer Cells by Targeting STAT4. J. Oncol. 2021, 2021, 5425491. [Google Scholar] [CrossRef] [PubMed]
  92. Otte, M.; Stachelscheid, J.; Glaß, M.; Wahnschaffe, L.; Jiang, Q.; Lone, W.; Ianevski, A.; Aittokallio, T.; Iqbal, J.; Hallek, M.; et al. The miR-141/200c-STAT4 Axis Contributes to Leukemogenesis by Enhancing Cell Proliferation in T-PLL. Cancers 2023, 15, 2527. [Google Scholar] [CrossRef] [PubMed]
  93. Litvinov, I.; Cordeiro, B.; Fredholm, S.; Ødum, N.; Zargham, H.; Huang, Y.; Zhou, Y.; Pehr, K.; Kupper, T.S.; Woetmann, A.; et al. Analysis of STAT4 expression in cutaneous T-cell lymphoma (CTCL) patients and patient-derived cell lines. Cell Cycle 2014, 13, 2975–2982. [Google Scholar] [CrossRef] [PubMed]
  94. Mui, A.; Wakao, H.; O’Farrell, A.; Harada, N.; Miyajima, A. Interleukin-3, granulocyte-macrophage colony stimulating factor and interleukin-5 transduce signals through two STAT5 homologs. EMBO J. 1995, 14, 1166–1175. [Google Scholar] [CrossRef] [PubMed]
  95. Murray, P.J. The JAK-STAT Signaling Pathway: Input and Output Integration. J. Immunol. 2007, 178, 2623–2629. [Google Scholar] [CrossRef] [Green Version]
  96. Schindler, C.; Levy, D.E.; Decker, T. JAK-STAT Signaling: From Interferons to Cytokines. J. Biol. Chem. 2007, 282, 20059–20063. [Google Scholar] [CrossRef] [Green Version]
  97. Lin, J.-X.; Du, N.; Li, P.; Kazemian, M.; Gebregiorgis, T.; Spolski, R.; Leonard, W.J. Critical functions for STAT5 tetramers in the maturation and survival of natural killer cells. Nat. Commun. 2017, 8, 1320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Azam, M.; Erdjument-Bromage, H.; Kreider, B.; Xia, M.; Quelle, F.; Basu, R.; Saris, C.; Tempst, P.; Ihle, J.; Schindler, C. Interleukin-3 signals through multiple isoforms of Stat5. EMBO J. 1995, 14, 1402–1411. [Google Scholar] [CrossRef]
  99. Hosseini, R.; Sarvnaz, H.; Arabpour, M.; Ramshe, S.M.; Asef-Kabiri, L.; Yousefi, H.; Akbari, M.E.; Eskandari, N. Cancer exosomes and natural killer cells dysfunction: Biological roles, clinical significance and implications for immunotherapy. Mol. Cancer 2022, 21, 15. [Google Scholar] [CrossRef]
  100. Liu, C.; Yu, S.; Zinn, K.; Wang, J.; Zhang, L.; Jia, Y.; Kappes, J.C.; Barnes, S.; Kimberly, R.P.; Grizzle, W.E.; et al. Murine mammary carcinoma exosomes promote tumor growth by suppression of NK cell function. J. Immunol. 2006, 176, 1375–1385. [Google Scholar] [CrossRef]
  101. Marleau, A.M.; Chen, C.-S.; Joyce, J.A.; Tullis, R.H. Exosome removal as a therapeutic adjuvant in cancer. J. Transl. Med. 2012, 10, 134. [Google Scholar] [CrossRef] [Green Version]
  102. Catalano, M.; O’Driscoll, L. Inhibiting extracellular vesicles formation and release: A review of EV inhibitors. J. Extracell. Vesicles 2020, 9, 1703244. [Google Scholar] [CrossRef] [Green Version]
  103. Clayton, A.; Mitchell, J.P.; Court, J.; Linnane, S.; Mason, M.D.; Tabi, Z. Human Tumor-Derived Exosomes Down-Modulate NKG2D Expression. J. Immunol. 2008, 180, 7249–7258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Szczepanski, M.J.; Szajnik, M.; Welsh, A.; Whiteside, T.L.; Boyiadzis, M. Blast-derived microvesicles in sera from patients with acute myeloid leukemia suppress natural killer cell function via membrane-associated transforming growth factor-β1. Haematologica 2011, 96, 1302–1309. [Google Scholar] [CrossRef] [Green Version]
  105. Anton, O.M.; Peterson, M.E.; Hollander, M.J.; Dorward, D.W.; Arora, G.; Traba, J.; Rajagopalan, S.; Snapp, E.L.; Garcia, K.C.; Waldmann, T.A.; et al. Trans -endocytosis of intact IL-15Rα–IL-15 complex from presenting cells into NK cells favors signaling for proliferation. Proc. Natl. Acad. Sci. USA 2019, 117, 522–531. [Google Scholar] [CrossRef] [PubMed]
  106. Ye, S.-B.; Li, Z.-L.; Luo, D.-H.; Huang, B.-J.; Chen, Y.-S.; Zhang, X.-S.; Cui, J.; Zeng, Y.-X.; Li, J. Tumor-derived exosomes promote tumor progression and T-cell dysfunction through the regulation of enriched exosomal microRNAs in human nasopharyngeal carcinoma. Oncotarget 2014, 5, 5439–5452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Mantovani, A.; Sozzani, S.; Locati, M.; Allavena, P.; Sica, A. Macrophage polarization: Tumor-associated macrophages as a paradigm for polarized M2 mononuclear phagocytes. Trends Immunol. 2002, 23, 549–555. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, J.; Huang, H.; Lu, J.; Bi, P.; Wang, F.; Liu, X.; Li, X. Tumor cells induced-M2 macrophage favors accumulation of Treg in nasopharyngeal carcinoma. Int. J. Clin. Exp. Pathol. 2017, 10, 8389–8401. [Google Scholar]
  109. Zhu, J.; Paul, W.E. CD4 T cells: Fates, functions, and faults. Blood 2008, 112, 1557–1569. [Google Scholar] [CrossRef] [Green Version]
  110. Sharma, P.; Hu-Lieskovan, S.; Wargo, J.A.; Ribas, A. Primary, Adaptive, and Acquired Resistance to Cancer Immunotherapy. Cell 2017, 168, 707–723. [Google Scholar] [CrossRef] [Green Version]
  111. Pardoll, D.M. The blockade of immune checkpoints in cancer immunotherapy. Nat. Rev. Cancer 2012, 12, 252–264. [Google Scholar] [CrossRef] [Green Version]
  112. Msaouel, P.; Genovese, G.; Gao, J.; Sen, S.; Tannir, N.M. TAM kinase inhibition and immune checkpoint blockade– a winning combination in cancer treatment? Expert Opin. Ther. Targets 2020, 25, 141–151. [Google Scholar] [CrossRef]
  113. Cassetta, L.; Kitamura, T. Targeting Tumor-Associated Macrophages as a Potential Strategy to Enhance the Response to Immune Checkpoint Inhibitors. Front. Cell Dev. Biol. 2018, 6, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Mantovani, A.; Marchesi, F.; Malesci, A.; Laghi, L.; Allavena, P. Tumour-associated macrophages as treatment targets in oncology. Nat. Rev. Clin. Oncol. 2017, 14, 399–416. [Google Scholar] [CrossRef] [PubMed]
  115. Ito, S.-E.; Shirota, H.; Kasahara, Y.; Saijo, K.; Ishioka, C. IL-4 blockade alters the tumor microenvironment and augments the response to cancer immunotherapy in a mouse model. Cancer Immunol. Immunother. 2017, 66, 1485–1496. [Google Scholar] [CrossRef]
  116. Wang, H.-W.; Joyce, J.A. Alternative activation of tumor-associated macrophages by IL-4. Cell Cycle 2010, 9, 4824–4835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Binnemars-Postma, K.; Bansal, R.; Storm, G.; Prakash, J. Targeting the Stat6 pathway in tumor-associated macrophages reduces tumor growth and metastatic niche formation in breast cancer. FASEB J. 2018, 32, 969–978. [Google Scholar] [CrossRef] [Green Version]
  118. He, K.; Barsoumian, H.B.; Puebla-Osorio, N.; Hu, Y.; Sezen, D.; Wasley, M.D.; Bertolet, G.; Zhang, J.; Leuschner, C.; Yang, L.; et al. Inhibition of STAT6 with Antisense Oligonucleotides Enhances the Systemic Antitumor Effects of Radiotherapy and Anti–PD-1 in Metastatic Non–Small Cell Lung Cancer. Cancer Immunol. Res. 2023, 11, 486–500. [Google Scholar] [CrossRef]
  119. Kamerkar, S.; Leng, C.; Burenkova, O.; Jang, S.C.; McCoy, C.; Zhang, K.; Dooley, K.; Kasera, S.; Zi, T.; Sisó, S.; et al. Exosome-mediated genetic reprogramming of tumor-associated macrophages by exoASO-STAT6 leads to potent monotherapy antitumor activity. Sci. Adv. 2022, 8, eabj7002. [Google Scholar] [CrossRef]
  120. Kong, W.; Zhang, L.; Chen, Y.; Yu, Z.; Zhao, Z. Cancer cell-derived exosomal LINC00313 induces M2 macrophage differentiation in non-small cell lung cancer. Clin. Transl. Oncol. 2022, 24, 2395–2408. [Google Scholar] [CrossRef]
  121. Park, J.E.; Dutta, B.; Tse, S.W.; Gupta, N.; Tan, C.F.; Low, J.K.; Yeoh, K.W.; Kon, O.L.; Tam, J.P.; Sze, S.K. Hypoxia-induced tumor exosomes promote M2-like macrophage polarization of infiltrating myeloid cells and microRNA-mediated metabolic shift. Oncogene 2019, 38, 5158–5173. [Google Scholar] [CrossRef] [PubMed]
  122. Peng, Y.; Zhao, M.; Hu, Y.; Guo, H.; Zhang, Y.; Huang, Y.; Zhao, L.; Chai, Y.; Wang, Z. Blockade of exosome generation by GW4869 inhibits the education of M2 macrophages in prostate cancer. BMC Immunol. 2022, 23, 37. [Google Scholar] [CrossRef] [PubMed]
  123. Bellmunt, M.; López-Puerto, L.; Lorente, J.; Closa, D. Involvement of extracellular vesicles in the macrophage-tumor cell communication in head and neck squamous cell carcinoma. PLoS ONE 2019, 14, e0224710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Sica, A.; Mantovani, A. Macrophage plasticity and polarization: In vivo veritas. J. Clin. Investig. 2012, 122, 787–795. [Google Scholar] [CrossRef]
  125. Pauleau, A.-L.; Rutschman, R.; Lang, R.; Pernis, A.; Watowich, S.S.; Murray, P.J. Enhancer-Mediated Control of Macrophage-Specific Arginase I Expression. J. Immunol. 2004, 172, 7565–7573. [Google Scholar] [CrossRef] [PubMed]
  126. Rodig, S.J.; Meraz, M.A.; White, J.; A Lampe, P.; Riley, J.K.; Arthur, C.D.; King, K.L.; Sheehan, K.C.; Yin, L.; Pennica, D.; et al. Disruption of the Jak1 Gene Demonstrates Obligatory and Nonredundant Roles of the Jaks in Cytokine-Induced Biologic Responses. Cell 1998, 93, 373–383. [Google Scholar] [CrossRef] [Green Version]
  127. O’Shea, D.M.S.J.J.; Schwartz, D.M.; Villarino, A.V.; Gadina, M.; McInnes, I.B.; Laurence, A. The JAK-STAT Pathway: Impact on Human Disease and Therapeutic Intervention. Annu. Rev. Med. 2015, 66, 311–328. [Google Scholar] [CrossRef] [Green Version]
  128. Ihle, J.N.; A Witthuhn, B.; Quelle, F.W.; Yamamoto, K.; Silvennoinen, O. Signaling Through the Hematopoietic Cytokine Receptors. Annu. Rev. Immunol. 1995, 13, 369–398. [Google Scholar] [CrossRef]
  129. Martinez-Cardús, A.; Martinez-Balibrea, E.; Bandrés, E.; Malumbres, R.; Ginés, A.; Manzano, J.L.; Taron, M.; Garcia-Foncillas, J.; Abad, A. Pharmacogenomic approach for the identification of novel determinants of acquired resistance to oxaliplatin in colorectal cancer. Mol. Cancer Ther. 2009, 8, 194–202. [Google Scholar] [CrossRef] [Green Version]
  130. Rodriguez, C.; Hughes-Davies, L.; Vallès, H.; Orsetti, B.; Cuny, M.; Ursule, L.; Kouzarides, T.; Theillet, C.; Finkle, D.; Quan, Z.R.; et al. Amplification of the BRCA2 Pathway Gene EMSY in Sporadic Breast Cancer Is Related to Negative Outcome. Clin. Cancer Res. 2004, 10, 5785–5791. [Google Scholar] [CrossRef] [Green Version]
  131. Edwards, J.; Krishna, N.S.; Witton, C.J.; Bartlett, J.M.S. Gene amplifications associated with the development of hormone-resistant prostate cancer. Clin. Cancer Res. 2003, 9, 5271–5281. [Google Scholar]
  132. Tang, L.-Y.; Heller, M.; Meng, Z.; Yu, L.-R.; Tang, Y.; Zhou, M.; Zhang, Y.E. Transforming Growth Factor-β (TGF-β) Directly Activates the JAK1-STAT3 Axis to Induce Hepatic Fibrosis in Coordination with the SMAD Pathway. J. Biol. Chem. 2017, 292, 4302–4312. [Google Scholar] [CrossRef] [Green Version]
  133. Xing, H.; Liang, C.; Xu, X.; Sun, H.; Ma, X.; Jiang, Z. Mesenchymal stroma/stem-like cells of GARP knockdown inhibits cell proliferation and invasion of mouse colon cancer cells (MC38) through exosomes. J. Cell. Mol. Med. 2020, 24, 13984–13990. [Google Scholar] [CrossRef] [PubMed]
  134. Jiang, M.; Ye, J.; Wang, X.; Li, N.; Wang, Y.; Shi, Y. Phosphatase SHP1 impedes mesenchymal stromal cell immunosuppressive capacity modulated by JAK1/STAT3 and P38 signals. Cell Biosci. 2020, 10, 65. [Google Scholar] [CrossRef]
  135. Li, W.; Rosenbruch, M.; Pauluhn, J. Effect of PEEP on phosgene-induced lung edema: Pilot study on dogs using protective ventilation strategies. Exp. Toxicol. Pathol. 2015, 67, 109–116. [Google Scholar] [CrossRef] [PubMed]
  136. Luo, S.; Trübel, H.; Wang, C.; Pauluhn, J. Phosgene- and chlorine-induced acute lung injury in rats: Comparison of cardiopulmonary function and biomarkers in exhaled breath. Toxicology 2014, 326, 109–118. [Google Scholar] [CrossRef] [PubMed]
  137. Fan, E.; Brodie, D.; Slutsky, A.S. Acute Respiratory Distress Syndrome. JAMA 2018, 319, 698–710. [Google Scholar] [CrossRef]
  138. Jiang, Z.; Shao, Y.; Zhang, L.; Shen, J. Lung-derived exosomes regulate the function of mesenchymal stem cells and alleviate phosgene-induced lung injury via miR-34c-3p. J. Biochem. Mol. Toxicol. 2021, 35, e22851. [Google Scholar] [CrossRef]
  139. Decker, T.; Müller, M. Jak-Stat Signaling: From Basics to Disease. Springer Science & Business Media. 2012. Available online: https://play.google.com/store/books/details?id=YBJGnD6rmR4C (accessed on 4 April 2023).
  140. Shuai, K. Modulation of STAT signaling by STAT-interacting proteins. Oncogene 2000, 19, 2638–2644. [Google Scholar] [CrossRef] [Green Version]
  141. Tanaka, T.; Soriano, M.A.; Grusby, M.J. SLIM Is a Nuclear Ubiquitin E3 Ligase that Negatively Regulates STAT Signaling. Immunity 2005, 22, 729–736. [Google Scholar] [CrossRef] [Green Version]
  142. Xue, C.; Xie, J.; Zhao, D.; Lin, S.; Zhou, T.; Shi, S.; Shao, X.; Lin, Y.; Zhu, B.; Cai, X. The JAK/STAT3 signalling pathway regulated angiogenesis in an endothelial cell/adipose-derived stromal cell co-culture, 3D gel model. Cell Prolif. 2016, 50, e12307. [Google Scholar] [CrossRef] [PubMed]
  143. Masoumi-Dehghi, S.; Babashah, S.; Sadeghizadeh, M. microRNA-141-3p-containing small extracellular vesicles derived from epithelial ovarian cancer cells promote endothelial cell angiogenesis through activating the JAK/STAT3 and NF-κB signaling pathways. J. Cell Commun. Signal. 2020, 14, 233–244. [Google Scholar] [CrossRef]
  144. Schindler, C.; Strehlow, I. Cytokines and STAT Signaling. Adv. Pharmacol. 1999, 47, 113–174. [Google Scholar] [CrossRef]
  145. Antunes, A.R.P.; Scheyltjens, I.; Lodi, F.; Messiaen, J.; Antoranz, A.; Duerinck, J.; Kancheva, D.; Martens, L.; De Vlaminck, K.; Van Hove, H.; et al. Single-cell profiling of myeloid cells in glioblastoma across species and disease stage reveals macrophage competition and specialization. Nat. Neurosci. 2021, 24, 595–610. [Google Scholar] [CrossRef] [PubMed]
  146. Li, M.; Xu, H.; Qi, Y.; Pan, Z.; Li, B.; Gao, Z.; Zhao, R.; Li, G. Tumor-derived exosomes deliver the tumor suppressor miR-3591-3p to induce M2 macrophage polarization and promote glioma progression. Oncogene 2022, 41, 4618–4632. [Google Scholar] [CrossRef] [PubMed]
  147. Xu, J.; Zhang, J.; Zhang, Z.; Gao, Z.; Qi, Y.; Qiu, W.; Pan, Z.; Guo, Q.; Li, B.; Zhao, S.; et al. Hypoxic glioma-derived exosomes promote M2-like macrophage polarization by enhancing autophagy induction. Cell Death Dis. 2021, 12, 373. [Google Scholar] [CrossRef] [PubMed]
  148. Pan, Z.; Zhao, R.; Li, B.; Qi, Y.; Qiu, W.; Guo, Q.; Zhang, S.; Zhao, S.; Xu, H.; Li, M.; et al. EWSR1-induced circNEIL3 promotes glioma progression and exosome-mediated macrophage immunosuppressive polarization via stabilizing IGF2BP3. Mol. Cancer 2022, 21, 16. [Google Scholar] [CrossRef] [PubMed]
  149. Zhang, Z.; Xu, J.; Chen, Z.; Wang, H.; Xue, H.; Yang, C.; Guo, Q.; Qi, Y.; Guo, X.; Qian, M.; et al. Transfer of MicroRNA via Macrophage-Derived Extracellular Vesicles Promotes Proneural-to-Mesenchymal Transition in Glioma Stem Cells. Cancer Immunol. Res. 2020, 8, 966–981. [Google Scholar] [CrossRef]
  150. Rane, S.G.; Reddy, E.P. JAK3: A novel JAK kinase associated with terminal differentiation of hematopoietic cells. Oncogene 1994, 9, 2415–2423. [Google Scholar]
  151. Saijo, K.; Park, S.Y.; Ishida, Y.; Arase, H.; Saito, T. Crucial Role of Jak3 in Negative Selection of Self-reactive T Cells. J. Exp. Med. 1997, 185, 351–356. [Google Scholar] [CrossRef] [Green Version]
  152. Purvis, I.J.; Avilala, J.; Guda, M.R.; Venkataraman, S.; Vibhakar, R.; Tsung, A.J.; Velpula, K.K.; Asuthkar, S. Role of MYC-miR-29-B7-H3 in Medulloblastoma Growth and Angiogenesis. J. Clin. Med. 2019, 8, 1158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. General constituents of the exosome. Exosomes contain proteins, endosome-specific proteins, cytoskeletal proteins, chaperones and enzymes, and various nucleic acids such as miRNA, PIWI-interacting RNA, circular RNA, and other non-coding RNA. Their membranes also contain abundant molecules such as integrins, tetraspanins, MHC I/II- major histocompatibility complex, MCT-1-Monocarboxylate transporter 1, and clusters of differentiation (CD) markers [4,5,8,9,10,11,12,13,14]. Created with BioRender.com.
Figure 1. General constituents of the exosome. Exosomes contain proteins, endosome-specific proteins, cytoskeletal proteins, chaperones and enzymes, and various nucleic acids such as miRNA, PIWI-interacting RNA, circular RNA, and other non-coding RNA. Their membranes also contain abundant molecules such as integrins, tetraspanins, MHC I/II- major histocompatibility complex, MCT-1-Monocarboxylate transporter 1, and clusters of differentiation (CD) markers [4,5,8,9,10,11,12,13,14]. Created with BioRender.com.
Pharmaceutics 15 01904 g001
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

Gombos, G.; Németh, N.; Pös, O.; Styk, J.; Buglyó, G.; Szemes, T.; Danihel, L.; Nagy, B.; Balogh, I.; Soltész, B. New Possible Ways to Use Exosomes in Diagnostics and Therapy via JAK/STAT Pathways. Pharmaceutics 2023, 15, 1904. https://doi.org/10.3390/pharmaceutics15071904

AMA Style

Gombos G, Németh N, Pös O, Styk J, Buglyó G, Szemes T, Danihel L, Nagy B, Balogh I, Soltész B. New Possible Ways to Use Exosomes in Diagnostics and Therapy via JAK/STAT Pathways. Pharmaceutics. 2023; 15(7):1904. https://doi.org/10.3390/pharmaceutics15071904

Chicago/Turabian Style

Gombos, Gréta, Nikolett Németh, Ondrej Pös, Jakub Styk, Gergely Buglyó, Tomas Szemes, Ludovit Danihel, Bálint Nagy, István Balogh, and Beáta Soltész. 2023. "New Possible Ways to Use Exosomes in Diagnostics and Therapy via JAK/STAT Pathways" Pharmaceutics 15, no. 7: 1904. https://doi.org/10.3390/pharmaceutics15071904

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