Publications
& News

Salivary Extracellular Vesicle-Associated exRNA as Cancer Biomarker.

1. January 2019

cancers
Review
Salivary Extracellular Vesicle-Associated exRNA as Cancer Biomarker
Giulia Chiabotto 1 , Chiara Gai 1, Maria Chiara Deregibus 2 and Giovanni Camussi 1,*
1 2
Received: 21 May 2019; Accepted: 22 June 2019; Published: 26 June 2019
Department of Medical Sciences, University of Torino, Torino 10126, Italy
2i3T Business Incubator and Technology Transfer, University of Torino, Torino 10126, Italy * Correspondence: giovanni.camussi@unito.it
Abstract: Extracellular vesicles (EVs) secreted in biological fluids contain several transcripts of the cell of origin, which may modify the functions and phenotype of proximal and distant cells. Cancer-derived EVs may promote a favorable microenvironment for cancer growth and invasion by acting on stroma and endothelial cells and may favor metastasis formation. The transcripts contained in cancer EVs may be exploited as biomarkers. Protein and extracellular RNA (exRNA) profiling in patient bio-fluids, such as blood and urine, was performed to identify molecular features with potential diagnostic and prognostic values. EVs are concentrated in saliva, and salivary EVs are particularly enriched in exRNAs. Several studies were focused on salivary EVs for the detection of biomarkers either of non-oral or oral cancers. The present paper provides an overview of the available studies on the diagnostic potential of exRNA profiling in salivary EVs.
Keywords: miRNA; non coding RNA; exosomes; microvesicles; cancer; saliva

  1. Introduction
    The aim of liquid biopsy is to identify biomarkers with diagnostic, predictive and prognostic values in bio-fluids, to avoid more invasive approaches. Researchers focused on different types of biomarkers, including proteins, circulating DNA fragments and cells, and extracellular RNAs (exRNAs). ExRNAs are more sensitive and specific biomarkers than proteins and better reflect the cell dynamic than DNA does [1]. However, several limitations in the use of exRNA as biomarkers still remain, related to their heterogeneity, the incomplete definition of their multiple targets and functions, and their stability in different biological fluids [2].
    Nowadays, the recently developed techniques of sequencing allow for an accurate evaluation of RNA expression, which reflects cellular genetic and functional states. Different types of RNA biomarkers have been considered in cancer. Differential mRNA expression profiles may reflect the positive and negative regulation of tumor-associated genes in several cancers and may provide suitable biomarkers for monitoring the clinical outcome of patients [3–5]. Non-coding RNAs, such as microRNAs (miRNAs), piwi-interacting RNA (piRNA), small nucleolar RNA (snoRNA), circular RNA (circRNA) and long non-coding RNAs (lncRNAs), have also been investigated as potential biomarkers in cancer [1]. Moreover, the detection of chimeric RNAs may allow for the identification of chromosomal aberrations [6,7]. The stability of different exRNAs in the biological fluids depends on protection from exonucleases, provided either by RNA binding proteins, such as those of the Argonaute family, and high- and low-density lipoproteins, and by encapsulation in membrane vesicles [8–10].
    Membrane vesicles released by cells in the extracellular space have recently emerged as a good evolutionarily preserved mechanism of inter-cellular communication. The vesicles are able to share genetic information among cells by delivering proteins, bio-active lipids and nucleic acids protected from degrading enzymes [11–13]. These vesicles, termed extracellular vesicles (EVs), are abundant
    􏰀􏰁􏰂􏰀􏰃 􏰅􏰆􏰇
    􏰈􏰉􏰊􏰋􏰌􏰂􏰍
    Cancers 2019, 11, 891; doi:10.3390/cancers11070891 www.mdpi.com/journal/cancers

Cancers 2019, 11, 891 2 of 21
in all biological fluids and can be exploited for searching biomarkers since they retain the molecular signature of the cell of origin.
One challenge of liquid biopsy is the choice of the bio-fluid that better reflects the occurrence of cancer. Most studies have focused on blood, but several other bio-fluids are now gaining attention, including saliva. Saliva is enriched in EVs and may represent a bio-fluid suitable for searching for markers of oral and systemic diseases.

  1. EVs as Carriers of exRNA
    EVs are a heterogeneous population, which includes membrane vesicles of different sizes and biogenesis. The three main categories of EVs include exosomes, ectosomes, and apoptotic bodies [14–17]. Exosomes are nano-sized vesicles (35–100 nm), which originate from the multivesicular bodies and are secreted by a process of exocytosis. This process requires the inward budding of multivesicular bodies-membrane, followed by fusion with plasma membrane and release in the extracellular space. The endosomal sorting complex required for transport (ESCRT) machinery, and several components of the Ras GTPases (RAB) family [18,19] and of the tetraspanin family [18], are involved in such processes. Vesicles generated by the budding of surface plasma membrane with the inclusion of cytoplasmic constituents have been termed microvesicles. This term is misleading as these vesicles include a large population of vesicles within the nano-range (60–250 nm), such as those released from healthy cells. It has been therefore suggested that one should name these vesicles ectosomes or shedding vesicles [14]. Shedding vesicles also include larger vesicles that may reach 1000 nm, and some of them may derive from cells in a pre-apoptotic phase. Microvesicle formation is related to the modification of plasma membrane curvature due to changes in lipid and protein interactions involving the arrestin domain-containing protein-1 (ARRDC1) and the late endosomal protein tumor susceptibility gene 101 (TSG101). The cytoskeleton rearrangements controlled by the signaling cascade of Ras-related GTPase ADP-ribosylation factor 6 (ARF6) promote vesiculation and release [20]. The apoptotic bodies released by cells undergoing programmed death are vesicles with a diameter of 1000–5000 nm and may contain nuclear fragments and intact chromosomes [21].
    Most of the studies on the use of EVs as potential biomarkers have been performed on exosomes and microvesicles, as both types of vesicles may encapsulate fragments of genomic and mitochondrial DNA origin [22–25], and different classes of RNA, such as mRNA, miRNA, lncRNA, mitochondrial RNA, transfer RNA, and ribosomal RNA [26–29]. Furthermore, nano-sized vesicles may be released by the same cell by exocytosis or by surface membrane budding, and it may be difficult to discriminate vesicles discharged by non-apoptotic cells on the basis of mechanisms of origin. In fact, some molecules constituent of the endosomal sorting complex required for transport (ESCRT) and some ancillary proteins such as TSG101, Alix and Vacuolar protein sorting-associated protein 4 (VPS4) implicated in the formation of exosomes, are also reported in the literature to also be shared by shedding vesicles [30]. The discharge of exosomes may involve some constituents of the RAB family of GTPase proteins implicated in the MVBs/plasma membrane interaction [31,32]. Furthermore, the biogenesis of shedding vesicles may depend on a reorganization of the proteins of the cytoskeleton myosin and actin under the control of the ARF6 signaling [30]. Some tetraspanins and some ESCRT proteins are often reported as common exosome and shedding vesicle markers and cannot represent a peculiarity principle [33]. However, CD9, CD63 and CD81 tetraspanins are reported to be enhanced in exosomes [18], while annexin A1 is considered a marker for microvesicles [34]. Due to the heterogeneity of EVs produced by different cell types and present in the biological fluids, the protocols used for EV purification assume a critical relevance. For this reason, the public available databases [35–37] take into account the procedures used for the purification of EVs when describing the lipid, protein and nucleic acid composition. Of interest, the comparative lipidomic, proteomic and genomic analyses between the cells of origin and their released EVs highlight the presence of qualitative and quantitative differences, in both basal and stimulated condition. These data suggest that the EV cargo is actively modulated [38–40]. The EV mediated transfer of their cargo into recipient cells can induce epigenetic and functional

Cancers 2019, 11, 891 3 of 21
changes into the recipient cells [41]. Some studies indicate that genetic materials encapsulated in EV include mitochondrial [24,42] and genomic [22] DNAs. Other studies performed on exosomal sub-fractions of EVs suggest that the DNA release is not related to the small vesicle release but to the autophagy- and multivesicular-endosome-dependent mechanism [34]. Several RNA species were found to be associated with EVs. EVs contain intact mRNA that can be translated into proteins in the recipient cells [27,28], but also many fragments of 200 nucleotides [43] that may have a biological role as scavengers and/or values as biomarkers. The exRNA enriched in EVs include miRNAs, ribosomal RNAs, tRNA fragments piRNA, snoRNA, Y-RNA, circRNA and lncRNAs [44–48].
Little is known about the process of nucleic acid compartimentalization into EVs [39,49–52]. Some proteins involved in EV biogenesis are potential candidates for RNA encapsulation in EVs. For instance, it has been shown that in EVs purified by differential ultracentrifugation from liver stem cells, Alix coprecipitate with Argonate 2 (Ago2) protein and miRNAs. The significant reduction of EV-associated miRNAs in Alix knock-down cells suggests that Alix can have a role as a component of ESCRT in the export of the Ago2-miRNA complex [53]. Through a high-resolution density gradient fractionation coupled with an immunoaffinity capture of exosomes, Argonaute proteins were detected in the non-vesicular compartment [34], which may contain components of the multivesicular body membranes. In breast cancer-derived EVs, miRNAs associated with Ago2 were shown to induce an alteration in the transcriptome of the recipient cells [54]. By regulating the Ago2 secretion [55], GTPase KRas (KRAS) has been involved in the miRNA compartmentalization into EVs released by colorectal cancer cells [56]. Moreover, miRNA packing into EVs depends on the interaction with the heterogeneous nuclear ribonucleoprotein A2B1 (hnRNPA2B1) [57] and with the RNA-binding protein Y-box protein I (YBX1) [58].

  1. EVs in Cancer Biology
    EVs released by cancer cells may act both locally, contributing to create a favorable microenvironment for cancer growth, and at distance, promoting the metastatic niche formation. Several studies have shown that cancer EVs contribute to the induction of chemo-resistance [59–63], to the remodeling of extracellular matrix [64,65], to tumor vascularization [66] and to epithelial-mesenchymal transition with a consequent enhanced migration/invasion and metastasis formation [67–69]. EVs also participate as active players in the bi-directional crosstalk between cancer cells and cells present in the microenvironment, such as fibroblasts [70,71], which may secrete EVs conferring chemo-resistance [72–74] and invasiveness to cancer cells [70,71].
    Several mechanisms of action involving the EV-mediated transfer of proteins and exRNA have been described and exploited as diagnostic markers. In particular, the miRNA-mediated effects have been extensively studied. Several miRNAs present in EVs released from breast cancer (miR-100, miR-222, miR30a and miR-17), lung cancer (miR-100-5p), and ovarian cancer (miR-21) or released from stromal cells (miR-21 and miR-146a) were shown to confer chemo-resistance [60–62,73,74] (Table 1). Cancer EVs may contribute to new blood vessel formation by transferring to recipient fibroblasts and endothelial cells pro-angiogenic miRNAs such as miR-155, miR-210 and miR-494, which are under the regulation of the hypoxia-inducible factor (HIF) 1α [75–79] (Table 2). Moreover, several studies on EVs released by cancer cells indicate that they promote the development of a pre-metastatic niche by transferring either proteins or oncogenic miRNAs [80–83]. For instance, some miRNAs (miR-125b, miR-130b and miR-155) present in prostate cancer and released by EVs have been shown to confer a protumorigenic phenotype to adipose-derived mesenchymal stem cells [84].

Cancers 2019, 11, 891 4 of 21
Table 1. The role of EVs in chemo-resistance and immune-modulation. Several proteins and exRNAs have been described to be involved in tumor chemo-resistance and immune-modulation.
Biological Effect
Resistance to chemotherapy
Mechanism of Action
Transfer of MDR-1/P-gp
Transfer of miR-100, miR-222, miR-30a and miR-17
Transfer of miR-21
Transfer of miR-100-5p, miR-21 and miR-133b
Transfer of miR-21, which downregulates APAF1
Transfer of miR-146a with Snail mRNA
Activation of the antiviral/ NOTCH3 signaling pathway
Release of pro-inflammatory cytokines by macrophages, possibly mediated by miR-21 and miR-29a
Inhibition of dendritic cell maturation and functions, by delivering specific miRNAs (e.g., miR-203, miR-212-3p)
MDSCs activation, which leads to TGF-β-mediated suppression of T cell activity
Suppression of the T-cell activity mediated by PDL-1, TGF-β, Fas ligand and TRAIL
Inhibition of NK cell cytotoxic activity, possibly mediated by MIC A ligand of NKG2D receptor
Activation of a tumor antigen-specific immune response in humans
Cell Source
Docetaxel-resistant prostate cancer
Adriamycin and docetaxel-resistant breast cancer
Platinum-resistant ovarian cancer Cisplatin-resistant lung cancer Stroma Cancer-Associated Fibroblasts Stroma
Breast and lung cancer, melanoma
Renal carcinoma, pancreatic cancer, melanoma
Melanoma and colorectal carcinoma
Melanoma, colorectal, gastric and prostate cancer, head and neck squamous cell carcinoma
Mammary carcinoma, melanoma, cervical, head and neck, liver cancer
Melanoma and non-small cell lung cancer patients-derived dendritic cells
Target
Docetaxel-sensitive prostate cancer
Adriamycin and docetaxel-sensitive breast cancer
Platinum-sensitive ovarian cancer
Cisplatin-sensitive lung cancer
Ovarian cancer
Pancreatic cancer
Breast cancer
Tumor cells, fibroblasts, endothelial cells, and immune cells
Dendritic and T cells CD14+ monocytes
CD8+T cells NK cells
systemic administration
References
[59]
[60]
[61] [62,63] [74] [73] [72]
[85–87] [87–89] [90,91]
[92–96] [97–99]
[100,101]
Tumor immune-escape
Enhancement of immune response
EVs: extracellular vesicles, MDR: multidrug resistance protein, APAF1: apoptotic protease-activating factor 1, MDSCs: myeloid-derived suppressor cells, TGF-β: trasforming growth factor-β, PDL-1: programmed death-ligand 1, TRAIL: tumor necrosis factor-related apoptosis-inducing ligand, NK: natural killer, MIC: MHC class I–related chain, NKG2D: NKG2-D type II integral membrane protein.

Cancers 2019, 11, 891 5 of 21
Table 2. Role of EVs as biomarkers of tumor progression. Cancer-derived EV content has been proposed as a tumor biomarker and has been related to several processes involved in tumor aggressiveness.
Biological Effect Tumor biomarkers
Pro-angiogenic effect Decrease cell-to-cell
adhesion Increase in cell
migration/invasion Development of
premetastatic niche
Mechanism of Action
Transfer of miR-21, miR-141, miR-200a, miR-200b, miR-200c, miR-203, miR-205 and miR-214
Transfer of miR-17-3p, miR-21, miR-29a, miR-106a, miR-146 miR-155, miR-191, miR-192, miR-203, miR-205, miR-210, miR-212 and miR-214
Transfer of miR-18a, miR-221 and miR-224 Transfer of proangiogenic miRNAs, mostly regulated by
HIF-1α (miR-155-5p, miR-210 and miR-494)
Reduction of E-cadherin, let-7i and β-catenin expression, and increase of Snail1-2, Twist1-2, Sip1, vimentin, ZEB2 and N-cadherin expression, activation of MAPK pathway
Lipids and proteins (e.g., CD81)-dependent stimulation of the cancer cell motility via Wnt signaling
Delivery of TYRP2, VLA4, HSP70, an HSP90 isoform and the MET oncoprotein
Exosomal expression of tumor-specific integrin patterns
Delivery of MIF
Delivery of specific oncogenic miRNAs, e.g., miR-125b, miR-130b and miR-155, which induce a neoplastic reprogramming of recipient cells
Cell Source
Ovarian cancer
Lung cancer
Hepatocellular carcinoma
Melanoma, hepatocellular, lung and renal adenocarcinoma
Breast and bladder cancer, melanoma
Cancer Associated Fibroblasts
Melanoma
Osteosarcoma, rhabdomyosarcoma, Wilms tumor, skin and uveal melanoma, breast, colorectal, pancreatic and gastric cancer
Pancreatic ductal adenocarcinoma Prostate, renal cancer
Target
Serum
Serum
Serum
CAFs and endothelial cells
Mammary and urothelial cells epithelial cells, primary melanocytes
Melanoma, breast and prostate cancer
Bone marrow progenitor cell
Brain, lung and liver epithelium
Kupffer cell Adipose-derived stem
cells, lung epithelium
References
[102] [85,103]
[104] [75,77–79,81]
[67–69]
[70,71] [83]
[82]
[80]
[81,84]
HIF-1a: hypoxia inducible factor 1α, HSP90: heat shock protein 90, MET: hepatocyte growth factor receptor., TYRP2: tyrosinase-related protein-2, VLA4: very late antigen 4, HSP70: heat
shock protein 70, MIF: macrophage migration inhibitory factor.

Cancers 2019, 11, 891 6 of 21
A contribution in favoring the tumor immune-escape of EVs released by cancer cells has been also suggested [105,106] (Table 1). The mechanisms involved the activation of tumor-associated macrophages [85–87], suppressor myeloid cells [90,91] and the inhibition of NK cell activity [97–99]. By expressing PDL1 [92,93], the transforming growth factor (TGF) beta [94], the tumor necrosis factor–related apoptosis-inducing ligand (TRAIL) and the Fas ligand [95,96], cancer EVs exhibit an immunosuppressive activity on T cells. Moreover, cancer EVs inhibit the maturation of dendritic cells through a mechanism involving the expression of HLA-G [88] and specific miRNAs, such as miR-203 and miR-212-3p [89]. Despite tumor EVs have been mainly implicated in the tumor immune escape, they can also be exploited to cross-present tumor antigens to the antigen presenting cells eliciting an antigen-specific cytotoxic lymphocyte anti-tumoral response [100,101]. However, clinical trials based on this assumption have provided conflicting results [107–109].
At present, most of the studies looking for exRNAs as cancer biomarkers have been performed on whole blood, urine and cerebrospinal fluids. Recently, several studies have explored the detection of exRNAs associated with the EVs. Despite the fact that the quantitative and stoichiometric analyses revealed that many miRNAs are present in less than one single copy per single exosome [110], several studies indicate a potential utility as cancer biomarkers [102,103]. For instance, in hepatocellular carcinoma, the expression by serum EVs of miR-18a, miR-221and miR-224 has been suggested as potential diagnostic biomarkers [104]. Fabbri et al. demonstrated that EV-associated miR-21 and miR-29a bound to a Toll-like receptor family favoring an inflammatory pro-metastatic response in lung [85]. On the other hand, the expression of miR-21 in serum EVs in patients with breast cancer correlates with a favorable outcome [111]. By comparing the miRNA signature of ovarian cancer EVs with that of EVs from normal subjects, Taylor and colleagues suggested a potential utility to screening asymptomatic patients [102]. A significant similarity of EV-associated miRNAs was observed with tumor-derived miRNAs in lung adenocarcinoma [103]. Moreover, the miRNA patterns of patients were clearly distinct from those of normal controls, suggesting that circulating EV-associated miRNAs might be useful as a non-invasive screening test [103].

  1. Salivary EVs as Biomarkers
    EVs are particularly enriched in saliva, which in respect to blood does not undergo coagulation. This is an important issue because many studies have been performed on serum. Coagulation induces a consistent release of EVs from platelets, thus modifying the composition of circulating EVs [112]. Salivary EVs should derive in part from salivary glands and in part from circulation: indeed, about a 30% similarity of salivary and plasma proteome has been described by a few studies [113–115]. In particular, using liquid chromatography and mass spectrometry, 19,474 unique peptides have been isolated from whole saliva in a multicenter study [113]. Protein annotation was assessed by matching the identified peptides with a recently published dataset of the human plasma proteome [116], and 1939 different proteins were identified as commonly expressed in blood and saliva. However, a puzzling aspect is the expression of neuronal markers in salivary EVs with significant changes in the miRNA pattern and in the proteomic profile after a head concussion [117] and in neurological diseases [118,119]. Moreover, the EV composition may be affected by the presence in saliva of viruses, including the human papillomavirus (HPV) [120–124] and the neurotropic human herpesviruses (e.g., HHV-6), which are detectable in the saliva of infected subjects [125].
    A critical aspect in the use of salivary EVs as biomarkers is the purification technique that is used (Table 3). In fact, results may vary depending on the purified subpopulations and the presence of contaminants, such as bacterial flora. Therefore, accurate mouth washing, careful standardization on saliva collection and sample filtration are recommended to abate the bacterial load.

Cancers 2019, 11, 891 7 of 21
Table 3. Biomarkers detected in salivary EVs. Salivary EVs can be purified using different EV isolation techniques and can be exploited as biomarkers because they contain disease-related proteins and exRNA.
Disease
Brain injury and neurological disorders
Oral squamous cell carcinoma
Lung cancer
Head and neck carcinoma
Pancreatic cancer
Isolation Method
Differential ultracentrifugation XYCQ EV Enrichment KIT Differential ultracentrifugation Differential ultracentrifugation
Charge-based precipitation
Affinity chromatography column combined with filter system (ACCF)
Affinity chromatography column combined with filter system (ACCF)
Differential ultracentrifugation Total Exosome Isolation Reagent
(Invitrogen) Differential ultracentrifugation
EV Biomarkers
CDC2, CSNK1A1, and CTSD
α-synuclein
CD63
PPIA
miR-412-3p, miR-512-3p, miR-27a-3p, miR-494-3p, miR-302b-3p, miR-517b-3p
Annexin A1, A2, A3, A5, A6, A11; NPRL2; CEACAM1; MUC1; PROM1; HIST1H4A; TNFAIP3
BPIFA1, CRNN, MUC5B, IQGAP
miR-486-5p, miR-486-3p, miR-10b-5p, miR-122
miR-1246, miR-4644
Apbb1ip, Aspn, BCO31781, Daf2, Foxp1, Gng2, Incenp
Type of Biomarker
mRNA protein protein protein
miRNA
protein
protein miRNA miRNA mRNA
References
[117]
[119] [126,127] [128]
[129] [130]
[131]
[132]
[133]
[134]
CDC2: Cyclin-dependent kinase A-1, CSNK1A1: Casein Kinase 1 Alpha 1, CTSD: Cathepsin D, PPIA: Peptidyl-prolyl cis-trans isomerase A, NPRL2: GATOR complex protein NPRL2, CEACAM1: Carcinoembryonic antigen-related cell adhesion molecule 1, MUC1: Mucin 1, PROM1: Prominin 1, HIST1H4A: Histone H4, TNFAIP3: Tumor necrosis factor alpha-induced protein 3, BPIFA1: BPI fold-containing family A member 1, CRNN: Cornulin, MUC5B: Mucin 5b, IQGAP: Ras GTPase-activating-like protein IQGAP1.

Cancers 2019, 11, 891 8 of 21
Differential ultracentrifugation or density gradient ultracentrifugation are considered the gold standard for the purification of EV subpopulations. These techniques have been further implemented with the combined use of the immune-affinity capture of exosomes [34]. To improve the separation of vesicles from non-vesicular components a floating technique has been proposed, based on gradient fractioning centrifugation, with samples applied to the bottom of tubes [135]. However, the standardization of these techniques may be difficult, as the results are influenced not only by the centrifugal radius of the rotor and g force type, but also by the viscosity of the starting solution. In addition, due to mechanical damage, membrane debris are generated, as seen by electron microscopy. Moreover, the difficult detection of proteins and RNAs has been described [136–139]. To avoid shear stress due to ultracentrifugation, size exclusion chromatography has been employed with the aim to separate small vesicles from protein contaminants [140–142]. Immuno-affinity purification allows for the recovery of sub fractions of EVs based on the expression of surface markers [139,143–145], and several kits are commercially available. Microfiltration has also been used with membranes with appropriate pore sizes to remove cell debris and apoptotic bodies [143]. However, this technique is limited by EV adhesion to membranes and pore clogging. In addition, to isolate small biological samples, all these techniques may have a low efficient recovery of EVs. Another approach for isolating EVs from biological liquids is based on polymeric precipitation [146–151]. This approach allows for a rapid precipitation of EVs, but it is limited by the co-precipitation of proteins of a non-vesicular origin such as lipoproteins [136,152,153]. Recently, a new technique based on electric field-induced release and measurement has been successfully applied to liquid biopsy in saliva [154]. Using this technique, the mutation of epidermal growth factor receptor (EGFR) in patients with lung cancers was detected and matched with biopsy genotyping [154,155]. Moreover, the electric field-induced release has been combined with the magnetic beads immune-capturing of exosomes [156,157], resulting in a highly sensitive and specific method of exRNA extraction and analysis. Compared to polymeric precipitation and differential centrifugation, this approach is less time consuming, requires smaller sample volumes and does not involve sample lysis that may reduce exRNA yield. However, for each EV extraction, the capture probe that is attached to the magnetic beads allows for the isolation of only those EVs containing the exosome-specific surface marker used for capturing EVs [156]. In fact, EVs are a heterogeneous population of vesicles, and individual EV analyses show that not all EVs co-express the same tetraspanin. Therefore, this technique may not include the whole pattern of EV-associated exRNA.
By quantitative nano-structural and single molecule force spectroscopy, Sharma et al. [126] performed a bio-molecular analysis of exosomes present in the saliva from patients with oral cancer. They demonstrated that exosomes were augmented in number and size, displayed a dissimilar morphology and showed an increased expression of CD63. Similarly, Zlotogorski-Hurvitz et al. [127] described a bigger salivary exosome concentration and size in patients with oral cancers in comparison with healthy subjects, a higher expression of CD63 and a decreased expression of CD9 and CD81. Few other studies performed a proteomic analysis of salivary exosomes in search of potential biomarkers of oral [128] and lung carcinomas [130]. A higher expression of the CD63 molecule was observed in EVs from the saliva of patients with oral cancers in respect to normal subjects [126]. Sun et al. performed a comparative proteomic analysis of salivary EVs in normal subjects and lung cancer patients [131]. In this study, several proteins were found to be dysregulated, and four of them were present in both salivary microvesicles and exosomes, suggesting their potential use for the detection of lung cancer.
It has been reported that in saliva, the bulk of miRNAs is packaged in exosomes [13]. In fact, miRNAs are easily detectable in EVs present in saliva [158,159]. Several studies focused on the possibility of exRNA isolation from saliva and oral samples [160–162] and in particular on salivary EV associated miRNAs in patients with oral cancer [160–164].
Langevin et al. performed a comprehensive miRNA sequence analysis of EVs derived from the saliva of patients with head and neck carcinomas and identified a distinct pattern of secretion and, in particular, miRNAs secreted only by cancer cells [132]. Some miRNAs, such as miR-486-5p,

Cancers 2019, 11, 891 9 of 21
miR-486-3p and miR-10b-5p, were specifically overexpressed in the EVs of a subset of head and neck carcinomas. Machida and colleagues showed that miR-1246 and miR-4644 present in salivary EVs are potential biomarkers of cancers of the pancreato-biliary tract [133]. Taken together, these analyses may provide the bases for the development of new tumor biomarkers (Table 3).
A transcriptomic signature specific for pancreatic [165] and ovarian cancers [166] and proteomic signature modifications in lung cancer [167] have been described in whole saliva. Zhang et al. [165] demonstrated that the combination of KRAS, metyl CpG binding domain protein 3 like 2 (MBD3L2), acrosomal vesicle 1 (ACRV1), and dolichyl-phosphate mannosyltransferase subunit 1 (DPM1) mRNAs in saliva may differentiate patients with pancreatic carcinomas from patients with chronic pancreatitis and healthy subjects with a high sensitivity and specificity. Moreover, a transcriptomic analysis of salivary EVs by next generation sequencing showed the presence of many coding and non-coding RNAs, such as mRNAs for several proteins, miRNAs, snoRNAs, piRNAs, and lncRNAs [48,168]. Palanisamy et al. [169] found, in exosomes isolated from saliva, 509 mRNA transcripts, which once incorporated in keratinocytes were able to modify the protein expression in these cells. Moreover, exosomes from adenocarcinoma of the pancreatic ducts were able to modify the biology of exosomes derived from the salivary gland and induce changes in the salivary biomarker profiles [134]. Similarly, they showed an interaction between exosomes derived from the human metastatic mammary gland epithelial adenocarcinoma cell line MDA-MB-231 cells and exosomes derived from the human submandibular gland (HSG) cells. This interaction induced an activation of the HSG cell transcriptional machinery with an increase of total cellular RNA and transcriptomic and proteomic changes [170]. Salivary EVs derived from patients with pancreatic carcinoma were shown to inhibit NK cell activation, thus favoring tumor immune escape [171].
We analyzed the zeta potential of salivary EVs and, based on their negative charge, we developed a charge-based precipitation protocol. This technique allows for the efficient recovery of exRNA from a salivary EV population with a very homogeneous size and shape [159] (Figure 1). In a recent work [129], we used the charge-based precipitation method to isolate EVs from the saliva of patients with oral squamous cell carcinoma (OSCC) to investigate the presence of exRNAs suitable as biomarkers. Our aim was to assess whether this quick, simple and efficient technique could be useful for detecting exRNA in the salivary EVs of patients with OSCC. The diagnosis of OSCC is based on oral examination and histological analysis. However, the identification of salivary biomarkers may have potential prognostic and therapeutic values. To exclude misleading results due to a different exposition to risk factors, at the time of patients’ recruitment, subjects included in our study were checked for their habits regarding smoking and alcohol consumption. In fact, smoke and alcohol consumption has been described as potentially affecting the composition of EV-associated exRNA [172,173]. Therefore, patients and controls were matched to obtain a similar distribution of risk factors among the two groups to reduce this bias. Moreover, 5 out 21 patients were positive to the Human Papilloma Virus (HPV). To avoid the detection of exRNA of viral origin, we screened EVs for the presence of about 800 miRNAs with human-specific primers. Although HPV infection may alter the EV release and cargo, we did not observe any significant change in the size and concentration of EVs from HPV-positive patients compared to negative patients. A differential expression of the EV miRNA signature in OSCC cells infected or not by HPV has been previously shown [124]. In this study, the authors observed that HPV infected cells released EVs enriched with 14 miRNAs, whereas non-infected cells overexpressed 19 miRNAs. The cohorts of patients we studied were too small to draw any conclusions. However, we did not observe the differential expression of miRNAs, which has been previously described for EVs released by in vitro infected OSCC cells.

    Cancers 2019, 11, 891 10 of 21

Figure 1. Salivary EVs characterization. A representative transmission electron microscopy image of Article
EVs isolated by a charge-based precipitation method, showing a carpet of vesicles in the nano-range. In the inset, the bars indicate the size of the extracellular vesicles (EVs). The preparation was stained
Towards a Decolonial Narrative Ethics
with NanoVan (JEOL Jem-1010 electron microscope, original magnification ×75,000; inset ×150,000). Hille Haker
By comparing the miRNA expression of cancer patients and matched controls, we observed an
up-regulatioLnoyooflamUiRn-iv41er2s-i3typ,CmhicRa-g5o1,2C-3hpic,amgoi,RIL-2670a6-360p, aUnSdA;mhihRa-k4e9r4@-l3upc.iendup;aTtieel.n: t+s1-w77it3h-5o0r8a-2l3s6q8uamous
cell carcinoma. MiR-512-3p and miR-412-3p were also potentially sensitive and specific biomarkers,
Received: date; Accepted: date; Published: date
as indicated by the high AUC values (0.847 and 0.871 respectively, with p values < 0.02) and a maximum Youden’s Index. Interestingly, we also observed an exclusive expression of miR-302b-3p and
Abstract: This essay explores the contribution of two works of German literature to a decolonial miR-517b-3p in cancer EVs. Moreover, we performed a bio-informatic analysis to better understand
narrative ethics. It analyzes the structures of colonialism, taking narratives as a medium of and for
whether the tumor-enriched miRNAs could be functionally related to the tumor. We observed that
ethical reflection, and reinterprets the ethical concepts of recognition and responsibility. This essay
eight tumor-related pathways were potentially targeted by these miRNAs. In particular, miR-512-3p
examines two stories. Franz Kafka’s Report to an Academy (1917) addresses the biological racism of and miR-27a-3p may target 7 and 20 genes, respectively, of the ErbB signaling pathway, which is known
the German scientists around 1900, unmasking the racism that renders apes (or particular people)
to promote cell proliferation and survival in cancer [174] and is activated in oral carcinomas [175–177].
the pre-life of human beings (or particular human beings). It also demonstrates that the politics of
MiR-512-3p, miR-27a-3p, and miR-302b-3p could potentially target proteoglycan genes and CD44
recognition, based on conditional (mis-)recognition, must be replaced by an ethics of mutual
involved in c-Fos-mediated cell invasion and migration [178], ERK1/2 phosphorylation [179] and recognition. Uwe Timm’s Morenga (1978) uses the cross-reference of history and fiction as aesthetic
the phenotype of oral cancer stem cells [180]. Moreover, miR-512-3p, miR-412-3p, miR-27a-3p, and principle, narrating the history of the German genocide of the Nama and Herero people at the
miR-302b-3p reduced the expression of TGFβR2, frequently reduced in cancer and stroma cells in beginning of the 20th century. Intercultural understanding, the novel shows, is impossible when it
patients with oral squamous carcinomas [181]. Increased levels of the oncogenic miR-27a-3p has also
is based on the conditional, colonial (mis-)recognition that echoes Kafka’s unmasking; furthermore,
been detected in EVs obtained from the plasma of OSCC patients [182]. In this study, a comparable
the novel illuminates the interrelation of recognition and responsibility that requires not only an
miRNA signature was observed between plasma EVs and EVs released by OSCC cells in vitro.
aesthetic ethics of reading based on attentiveness and response but also a political ethics that
Recent studies have shown that EVs also contain lncRNAs [183]. The expression of lncRNAs has confronts the (German) readers as historically situated agents who must take responsibility for their
not been investigated in salivary EVs. However, salivary lncRNAs may represent a potential marker
past.
for OSSC [184]. In fact, a subset of lncRNAs was correlated with high metastatic OSCC. In particular, the lncRNA HOTAIR was found to be highly expressed in the saliva of patients with lymph node
Keywords: narrative ethics; recognition; responsibility; decoloniality; Kafka; Timm; racism; metastasis. Therefore, besides miRNAs, the search for lncRNAs in salivary EVs could be a valuable
genocide; German Empire
diagnostic and prognostic tool for OSCC.
Humanities 2019, 8, x; doi:FOR PEER REVIEW www.mdpi.com/journal/humanities

Cancers 2019, 11, 891 11 of 21

  1. Conclusions
    Taken together, these studies suggest that EVs derived from cancer cells may modulate the function and may induce epigenetic changes in neighboring or distant cells. These biological effects are related to the delivery of transcripts that are specific of the originator cells. Several studies have shown a prominent role of exRNAs associated with vesicles. Since EVs may retain the molecular signature of the cell of origin, it has been suggested that they are a potential diagnostic exploitation. The salivary EV composition may reflect the presence of local or systemic diseases and has been investigated as a potential biomarker for both oral and non-oral cancers. Changes in the molecular composition of the EVs of non-oral cancers may either depend on their derivation from blood (since salivary glands are vascularized) or be the consequence of phenotypic changes occurring in gland cells (as the results of the stimulation by circulating cancer EVs). However, so far, available studies are relatively few and include a low number of patients. Further studies are necessary to optimize the protocol of EV isolation from saliva in order to obtain reproducible results. Moreover, the use of the EV content as a biomarker should take into account that this may be influenced by a number of cancer-associated risk factors, such as viral infections, smoking, alcohol abuse, as well as a number of non-cancer-associated factors related to concomitant diseases. However, these limitations in the use of EVs as biomarkers are not restricted to saliva, but may influence EVs derived from any biological fluid. Since saliva is an easily obtainable non-invasive bio-fluid particularly enriched in EVs, it may represent a new approach for cancer biomarker discovery. However, to define whether salivary EVs have a real clinical diagnostic and prognostic potential would require comparative studies between EVs derived from tumor cells, blood and saliva, which are not at present available.
    Author Contributions: All authors equally contributed to the conceptualization of the article. The research of the pertinent literature was performed by M.C.D., G.C.; writing—original draft preparation, G.C.; review and editing, C.G., M.C.D., and G.C.
    Funding: This work was supported by Associazione Italiana per la Ricerca sul Cancro (AIRC).
    Conflicts of Interest: G.C. is a component of Scientific Advisory Board of Unicyte AG. The authors declare no
    conflict of interest.
    References
  2. Xi, X.; Li, T.; Huang, Y.; Sun, J.; Zhu, Y.; Yang, Y.; Lu, Z.J. RNA Biomarkers: Frontier of Precision Medicine for Cancer. NonCoding RNA 2017, 3, 9. [CrossRef] [PubMed]
  3. Lan, H.; Lu, H.; Wang, X.; Jin, H. MicroRNAs as Potential Biomarkers in Cancer: Opportunities and Challenges. Biomed. Res. Int. 2015, 2015, 125094. [CrossRef] [PubMed]
  4. Martinez-Ledesma, E.; Verhaak, R.G.W.; Treviño, V. Identification of a Multi-Cancer Gene Expression Biomarker for Cancer Clinical Outcomes Using a Network-Based Algorithm. Sci. Rep. 2015, 5, 11966. [CrossRef] [PubMed]
  5. Cancer Genome Atlas Network. Comprehensive Molecular Portraits of Human Breast Tumours. Nature 2012, 490, 61–70. [CrossRef] [PubMed]
  6. Cuzick, J.; Swanson, G.P.; Fisher, G.; Brothman, A.R.; Berney, D.M.; Reid, J.E.; Mesher, D.; Speights, V.O.; Stankiewicz, E.; Foster, C.S.; et al. Prognostic Value of an RNA Expression Signature Derived from Cell Cycle Proliferation Genes in Patients with Prostate Cancer: A Retrospective Study. Lancet Oncol. 2011, 12, 245–255. [CrossRef]
  7. Asmann, Y.W.; Necela, B.M.; Kalari, K.R.; Hossain, A.; Baker, T.R.; Carr, J.M.; Davis, C.; Getz, J.E.; Hostetter, G.; Li, X.; et al. Detection of Redundant Fusion Transcripts as Biomarkers or Disease-Specific Therapeutic Targets in Breast Cancer. Cancer Res. 2012, 72, 1921–1928. [CrossRef] [PubMed]
  8. Attard, G.; Clark, J.; Ambroisine, L.; Fisher, G.; Kovacs, G.; Flohr, P.; Berney, D.; Foster, C.S.; Fletcher, A.; Gerald, W.L.; et al. Duplication of the Fusion of TMPRSS2 to ERG Sequences Identifies Fatal Human Prostate Cancer. Oncogene 2008, 27, 253–263. [CrossRef] [PubMed]

Cancers 2019, 11, 891 12 of 21

  1. Arroyo, J.D.; Chevillet, J.R.; Kroh, E.M.; Ruf, I.K.; Pritchard, C.C.; Gibson, D.F.; Mitchell, P.S.; Bennett, C.F.; Pogosova-Agadjanyan, E.L.; Stirewalt, D.L.; et al. Argonaute2 Complexes Carry a Population of Circulating MicroRNAs Independent of Vesicles in Human Plasma. Proc. Natl. Acad. Sci. USA 2011, 108, 5003–5008. [CrossRef]
  2. Vickers, K.C.; Palmisano, B.T.; Shoucri, B.M.; Shamburek, R.D.; Remaley, A.T. MicroRNAs Are Transported in Plasma and Delivered to Recipient Cells by High-Density Lipoproteins. Nat. Cell Biol. 2011, 13, 423–433. [CrossRef]
  3. Guallar, D.; Wang, J. RNA-Binding Proteins in Pluripotency, Differentiation, and Reprogramming. Front. Biol. (Beijing) 2014, 9, 389–409. [CrossRef]
  4. Quesenberry, P.J.; Aliotta, J.M. Cellular Phenotype Switching and Microvesicles. Adv. Drug Deliv. Rev. 2010, 62, 1141–1148. [CrossRef] [PubMed]
  5. Chen, X.; Liang, H.; Zhang, J.; Zen, K.; Zhang, C.-Y. Secreted MicroRNAs: A New Form of Intercellular Communication. Trends Cell Biol. 2012, 22, 125–132. [CrossRef] [PubMed]
  6. Gallo, A.; Tandon, M.; Alevizos, I.; Illei, G.G. The Majority of MicroRNAs Detectable in Serum and Saliva Is Concentrated in Exosomes. PLoS ONE 2012, 7, e30679. [CrossRef] [PubMed]
  7. Cocucci, E.; Meldolesi, J. Ectosomes and Exosomes: Shedding the Confusion between Extracellular Vesicles. Trends Cell Biol. 2015, 25, 364–372. [CrossRef] [PubMed]
  8. Bobrie, A.; Colombo, M.; Krumeich, S.; Raposo, G.; Théry, C. Diverse Subpopulations of Vesicles Secreted by Different Intracellular Mechanisms Are Present in Exosome Preparations Obtained by Differential Ultracentrifugation. J. Extracell. Vesicles 2012, 1, 18397. [CrossRef] [PubMed]
  9. Ratajczak, M.Z.; Ratajczak, J. Extracellular Microvesicles as Game Changers in Better Understanding the Complexity of Cellular Interactions-From Bench to Clinical Applications. Am. J. Med. Sci. 2017, 354, 449–452. [CrossRef]
  10. Ratajczak, J.; Wysoczynski, M.; Hayek, F.; Janowska-Wieczorek, A.; Ratajczak, M.Z. Membrane-Derived Microvesicles: Important and Underappreciated Mediators of Cell-to-Cell Communication. Leukemia 2006, 20, 1487–1495. [CrossRef]
  11. Kowal, J.; Tkach, M.; Théry, C. Biogenesis and Secretion of Exosomes. Curr. Opin. Cell Biol. 2014, 29, 116–125. [CrossRef]
  12. Hyenne, V.; Apaydin, A.; Rodriguez, D.; Spiegelhalter, C.; Hoff-Yoessle, S.; Diem, M.; Tak, S.; Lefebvre, O.; Schwab, Y.; Goetz, J.G.; et al. RAL-1 Controls Multivesicular Body Biogenesis and Exosome Secretion. J. Cell Biol. 2015, 211, 27–37. [CrossRef]
  13. Tricarico, C.; Clancy, J.; D’Souza-Schorey, C. Biology and Biogenesis of Shed Microvesicles. Small Gtpases 2017, 8, 220–232. [CrossRef]
  14. Hristov, M.; Erl, W.; Linder, S.; Weber, P.C. Apoptotic Bodies from Endothelial Cells Enhance the Number and Initiate the Differentiation of Human Endothelial Progenitor Cells in Vitro. Blood 2004, 104, 2761–2766. [CrossRef] [PubMed]
  15. Balaj, L.; Lessard, R.; Dai, L.; Cho, Y.-J.; Pomeroy, S.L.; Breakefield, X.O.; Skog, J. Tumour Microvesicles Contain Retrotransposon Elements and Amplified Oncogene Sequences. Nat. Commun. 2011, 2, 180. [CrossRef] [PubMed]
  16. Lázaro-Ibáñez, E.; Sanz-Garcia, A.; Visakorpi, T.; Escobedo-Lucea, C.; Siljander, P.; Ayuso-Sacido, A.; Yliperttula, M. Different GDNA Content in the Subpopulations of Prostate Cancer Extracellular Vesicles: Apoptotic Bodies, Microvesicles, and Exosomes. Prostate 2014, 74, 1379–1390. [CrossRef] [PubMed]
  17. Guescini, M.; Genedani, S.; Stocchi, V.; Agnati, L.F. Astrocytes and Glioblastoma Cells Release Exosomes Carrying MtDNA. J. Neural Transm. (Vienna) 2010, 117, 1–4. [CrossRef] [PubMed]
  18. Sansone, P.; Savini, C.; Kurelac, I.; Chang, Q.; Amato, L.B.; Strillacci, A.; Stepanova, A.; Iommarini, L.; Mastroleo, C.; Daly, L.; et al. Packaging and Transfer of Mitochondrial DNA via Exosomes Regulate Escape from Dormancy in Hormonal Therapy-Resistant Breast Cancer. Proc. Natl. Acad. Sci. USA 2017, 114, E9066–E9075. [CrossRef] [PubMed]
  19. Ratajczak, J.; Miekus, K.; Kucia, M.; Zhang, J.; Reca, R.; Dvorak, P.; Ratajczak, M.Z. Embryonic Stem Cell-Derived Microvesicles Reprogram Hematopoietic Progenitors: Evidence for Horizontal Transfer of MRNA and Protein Delivery. Leukemia 2006, 20, 847–856. [CrossRef] [PubMed]

Cancers 2019, 11, 891 13 of 21

  1. Deregibus, M.C.; Cantaluppi, V.; Calogero, R.; Lo Iacono, M.; Tetta, C.; Biancone, L.; Bruno, S.; Bussolati, B.; Camussi, G. Endothelial Progenitor Cell Derived Microvesicles Activate an Angiogenic Program in Endothelial Cells by a Horizontal Transfer of MRNA. Blood 2007, 110, 2440–2448. [CrossRef] [PubMed]
  2. Valadi, H.; Ekström, K.; Bossios, A.; Sjöstrand, M.; Lee, J.J.; Lötvall, J.O. Exosome-Mediated Transfer of MRNAs and MicroRNAs Is a Novel Mechanism of Genetic Exchange between Cells. Nat. Cell Biol. 2007, 9, 654–659. [CrossRef] [PubMed]
  3. Fatima, F.; Nawaz, M. Vesiculated Long Non-Coding RNAs: Offshore Packages Deciphering Trans-Regulation between Cells, Cancer Progression and Resistance to Therapies. Noncoding RNA 2017, 3, 10. [CrossRef] [PubMed]
  4. van Niel, G.; D’Angelo, G.; Raposo, G. Shedding Light on the Cell Biology of Extracellular Vesicles. Nat. Rev. Mol. Cell Biol. 2018, 19, 213–228. [CrossRef] [PubMed]
  5. Baietti, M.F.; Zhang, Z.; Mortier, E.; Melchior, A.; Degeest, G.; Geeraerts, A.; Ivarsson, Y.; Depoortere, F.; Coomans, C.; Vermeiren, E.; et al. Syndecan-Syntenin-ALIX Regulates the Biogenesis of Exosomes. Nat. Cell Biol. 2012, 14, 677–685. [CrossRef] [PubMed]
  6. D’Souza-Schorey, C.; Chavrier, P. ARF Proteins: Roles in Membrane Traffic and Beyond. Nat. Rev. Mol. Cell Biol. 2006, 7, 347–358. [CrossRef] [PubMed]
  7. Nabhan, J.F.; Hu, R.; Oh, R.S.; Cohen, S.N.; Lu, Q. Formation and Release of Arrestin Domain-Containing Protein 1-Mediated Microvesicles (ARMMs) at Plasma Membrane by Recruitment of TSG101 Protein. Proc. Natl. Acad. Sci. USA 2012, 109, 4146–4151. [CrossRef]
  8. Jeppesen, D.K.; Fenix, A.M.; Franklin, J.L.; Higginbotham, J.N.; Zhang, Q.; Zimmerman, L.J.; Liebler, D.C.; Ping, J.; Liu, Q.; Evans, R.; et al. Reassessment of Exosome Composition. Cell 2019, 177, 428–445. [CrossRef] [PubMed]
  9. Kalra, H.; Simpson, R.J.; Ji, H.; Aikawa, E.; Altevogt, P.; Askenase, P.; Bond, V.C.; Borràs, F.E.; Breakefield, X.; Budnik, V.; et al. Vesiclepedia: A Compendium for Extracellular Vesicles with Continuous Community Annotation. PLoS Biol. 2012, 10, e1001450. [CrossRef] [PubMed]
  10. Kim, D.-K.; Kang, B.; Kim, O.Y.; Choi, D.-S.; Lee, J.; Kim, S.R.; Go, G.; Yoon, Y.J.; Kim, J.H.; Jang, S.C.; et al. EVpedia: An Integrated Database of High-Throughput Data for Systemic Analyses of Extracellular Vesicles. J. Extracell. Vesicles 2013, 2, 20384. [CrossRef] [PubMed]
  11. Mathivanan, S.; Simpson, R.J. ExoCarta: A Compendium of Exosomal Proteins and RNA. Proteomics 2009, 9, 4997–5000. [CrossRef] [PubMed]
  12. Li, Y.; Zheng, Q.; Bao, C.; Li, S.; Guo, W.; Zhao, J.; Chen, D.; Gu, J.; He, X.; Huang, S. Circular RNA Is Enriched and Stable in Exosomes: A Promising Biomarker for Cancer Diagnosis. Cell Res. 2015, 25, 981–984. [CrossRef]
  13. Li, C.C.Y.; Eaton, S.A.; Young, P.E.; Lee, M.; Shuttleworth, R.; Humphreys, D.T.; Grau, G.E.; Combes, V.; Bebawy, M.; Gong, J.; et al. Glioma Microvesicles Carry Selectively Packaged Coding and Non-Coding RNAs Which Alter Gene Expression in Recipient Cells. RNA Biol. 2013, 10, 1333–1344. [CrossRef]
  14. Villarroya-Beltri, C.; Baixauli, F.; Gutiérrez-Vázquez, C.; Sánchez-Madrid, F.; Mittelbrunn, M. Sorting It out: Regulation of Exosome Loading. Semin. Cancer Biol. 2014, 28, 3–13. [CrossRef]
  15. Quesenberry, P.J.; Aliotta, J.; Deregibus, M.C.; Camussi, G. Role of Extracellular RNA-Carrying Vesicles in Cell Differentiation and Reprogramming. Stem Cell Res. 2015, 6, 153. [CrossRef] [PubMed]
  16. Guescini, M.; Guidolin, D.; Vallorani, L.; Casadei, L.; Gioacchini, A.M.; Tibollo, P.; Battistelli, M.; Falcieri, E.; Battistin, L.; Agnati, L.F.; et al. C2C12 Myoblasts Release Micro-Vesicles Containing MtDNA and Proteins Involved in Signal Transduction. Exp. Cell Res. 2010, 316, 1977–1984. [CrossRef] [PubMed]
  17. Batagov, A.O.; Kurochkin, I.V. Exosomes Secreted by Human Cells Transport Largely MRNA Fragments That Are Enriched in the 3’-Untranslated Regions. Biol. Direct 2013, 8, 12. [CrossRef] [PubMed]
  18. Cheng, L.; Sun, X.; Scicluna, B.J.; Coleman, B.M.; Hill, A.F. Characterization and Deep Sequencing Analysis of Exosomal and Non-Exosomal MiRNA in Human Urine. Kidney Int. 2014, 86, 433–444. [CrossRef] [PubMed]
  19. Crescitelli, R.; Lässer, C.; Szabó, T.G.; Kittel, A.; Eldh, M.; Dianzani, I.; Buzás, E.I.; Lötvall, J. Distinct
    RNA Profiles in Subpopulations of Extracellular Vesicles: Apoptotic Bodies, Microvesicles and Exosomes.
    J. Extracell. Vesicles 2013, 2, 20677. [CrossRef] [PubMed]
  20. Hill, A.F.; Pegtel, D.M.; Lambertz, U.; Leonardi, T.; O’Driscoll, L.; Pluchino, S.; Ter-Ovanesyan, D.; Nolte-’t
    Hoen, E.N.M. ISEV Position Paper: Extracellular Vesicle RNA Analysis and Bioinformatics. J. Extracell. Vesicles 2013, 2, 22859. [CrossRef] [PubMed]

Cancers 2019, 11, 891 14 of 21

  1. Huang, X.; Yuan, T.; Tschannen, M.; Sun, Z.; Jacob, H.; Du, M.; Liang, M.; Dittmar, R.L.; Liu, Y.; Liang, M.; et al. Characterization of Human Plasma-Derived Exosomal RNAs by Deep Sequencing. BMC Genom. 2013, 14, 319. [CrossRef] [PubMed]
  2. Ogawa, Y.; Taketomi, Y.; Murakami, M.; Tsujimoto, M.; Yanoshita, R. Small RNA Transcriptomes of Two Types of Exosomes in Human Whole Saliva Determined by next Generation Sequencing. Biol. Pharm. Bull. 2013, 36, 66–75. [CrossRef]
  3. Abels, E.R.; Breakefield, X.O. Introduction to Extracellular Vesicles: Biogenesis, RNA Cargo Selection, Content, Release, and Uptake. Cell. Mol. Neurobiol. 2016, 36, 301–312. [CrossRef]
  4. Zhang, Y.; Liu, D.; Chen, X.; Li, J.; Li, L.; Bian, Z.; Sun, F.; Lu, J.; Yin, Y.; Cai, X.; et al. Secreted Monocytic MiR-150 Enhances Targeted Endothelial Cell Migration. Mol. Cell 2010, 39, 133–144. [CrossRef]
  5. Collino, F.; Deregibus, M.C.; Bruno, S.; Sterpone, L.; Aghemo, G.; Viltono, L.; Tetta, C.; Camussi, G. Microvesicles Derived from Adult Human Bone Marrow and Tissue Specific Mesenchymal Stem Cells Shuttle Selected Pattern of MiRNAs. PLoS ONE 2010, 5, e11803. [CrossRef] [PubMed]
  6. Goldie, B.J.; Dun, M.D.; Lin, M.; Smith, N.D.; Verrills, N.M.; Dayas, C.V.; Cairns, M.J. Activity-Associated MiRNA Are Packaged in Map1b-Enriched Exosomes Released from Depolarized Neurons. Nucleic Acids Res. 2014, 42, 9195–9208. [CrossRef] [PubMed]
  7. Iavello, A.; Frech, V.S.L.; Gai, C.; Deregibus, M.C.; Quesenberry, P.J.; Camussi, G. Role of Alix in MiRNA Packaging during Extracellular Vesicle Biogenesis. Int. J. Mol. Med. 2016, 37, 958–966. [CrossRef] [PubMed]
  8. Melo, S.A.; Sugimoto, H.; O’Connell, J.T.; Kato, N.; Villanueva, A.; Vidal, A.; Qiu, L.; Vitkin, E.; Perelman, L.T.; Melo, C.A.; et al. Cancer Exosomes Perform Cell-Independent MicroRNA Biogenesis and Promote
    Tumorigenesis. Cancer Cell 2014, 26, 707–721. [CrossRef] [PubMed]
  9. McKenzie, A.J.; Hoshino, D.; Hong, N.H.; Cha, D.J.; Franklin, J.L.; Coffey, R.J.; Patton, J.G.; Weaver, A.M.
    KRAS-MEK Signaling Controls Ago2 Sorting into Exosomes. Cell Rep. 2016, 15, 978–987. [CrossRef]
    [PubMed]
  10. Cha, D.J.; Franklin, J.L.; Dou, Y.; Liu, Q.; Higginbotham, J.N.; Demory Beckler, M.; Weaver, A.M.; Vickers, K.;
    Prasad, N.; Levy, S.; et al. KRAS-Dependent Sorting of MiRNA to Exosomes. eLife 2015, 4, e07197. [CrossRef]
    [PubMed]
  11. Villarroya-Beltri, C.; Gutiérrez-Vázquez, C.; Sánchez-Cabo, F.; Pérez-Hernández, D.; Vázquez, J.;
    Martin-Cofreces, N.; Martinez-Herrera, D.J.; Pascual-Montano, A.; Mittelbrunn, M.; Sánchez-Madrid, F. Sumoylated HnRNPA2B1 Controls the Sorting of MiRNAs into Exosomes through Binding to Specific Motifs. Nat. Commun. 2013, 4, 2980. [CrossRef]
  12. Shurtleff, M.J.; Temoche-Diaz, M.M.; Karfilis, K.V.; Ri, S.; Schekman, R. Y-Box Protein 1 Is Required to Sort MicroRNAs into Exosomes in Cells and in a Cell-Free Reaction. ELife 2016, 5, e19276. [CrossRef]
  13. Corcoran, C.; Rani, S.; O’Brien, K.; O’Neill, A.; Prencipe, M.; Sheikh, R.; Webb, G.; McDermott, R.; Watson, W.; Crown, J.; et al. Docetaxel-Resistance in Prostate Cancer: Evaluating Associated Phenotypic Changes and Potential for Resistance Transfer via Exosomes. PLoS ONE 2012, 7, e50999. [CrossRef]
  14. Chen, W.; Liu, X.; Lv, M.; Chen, L.; Zhao, J.; Zhong, S.; Ji, M.; Hu, Q.; Luo, Z.; Wu, J.; et al. Exosomes from Drug-Resistant Breast Cancer Cells Transmit Chemoresistance by a Horizontal Transfer of MicroRNAs. PLoS ONE 2014, 9, e95240. [CrossRef]
  15. Crow, J.; Atay, S.; Banskota, S.; Artale, B.; Schmitt, S.; Godwin, A.K. Exosomes as Mediators of Platinum Resistance in Ovarian Cancer. Oncotarget 2017, 8, 11917–11936. [CrossRef] [PubMed]
  16. Qin, X.; Yu, S.; Zhou, L.; Shi, M.; Hu, Y.; Xu, X.; Shen, B.; Liu, S.; Yan, D.; Feng, J. Cisplatin-Resistant Lung Cancer Cell-Derived Exosomes Increase Cisplatin Resistance of Recipient Cells in Exosomal MiR-100-5p-Dependent Manner. Int. J. Nanomed. 2017, 12, 3721–3733. [CrossRef] [PubMed]
  17. Xiao, X.; Yu, S.; Li, S.; Wu, J.; Ma, R.; Cao, H.; Zhu, Y.; Feng, J. Exosomes: Decreased Sensitivity of Lung Cancer A549 Cells to Cisplatin. PLoS ONE 2014, 9, e89534. [CrossRef] [PubMed]
  18. Nawaz, M.; Shah, N.; Zanetti, B.R.; Maugeri, M.; Silvestre, R.N.; Fatima, F.; Neder, L.; Valadi, H. Extracellular Vesicles and Matrix Remodeling Enzymes: The Emerging Roles in Extracellular Matrix Remodeling, Progression of Diseases and Tissue Repair. Cells 2018, 7, 167. [CrossRef] [PubMed]
  19. Sung, B.H.; Ketova, T.; Hoshino, D.; Zijlstra, A.; Weaver, A.M. Directional Cell Movement through Tissues Is Controlled by Exosome Secretion. Nat. Commun. 2015, 6, 7164. [CrossRef] [PubMed]
  20. Song, W.; Yan, D.; Wei, T.; Liu, Q.; Zhou, X.; Liu, J. Tumor-Derived Extracellular Vesicles in Angiogenesis. Biomed. Pharm. 2018, 102, 1203–1208. [CrossRef] [PubMed]

Cancers 2019, 11, 891 15 of 21

  1. Galindo-Hernandez, O.; Serna-Marquez, N.; Castillo-Sanchez, R.; Salazar, E.P. Extracellular Vesicles from MDA-MB-231 Breast Cancer Cells Stimulated with Linoleic Acid Promote an EMT-like Process in MCF10A Cells. Prostaglandins Leukot. Essent. Fat. Acids 2014, 91, 299–310. [CrossRef] [PubMed]
  2. Franzen, C.A.; Blackwell, R.H.; Todorovic, V.; Greco, K.A.; Foreman, K.E.; Flanigan, R.C.; Kuo, P.C.; Gupta, G.N. Urothelial Cells Undergo Epithelial-to-Mesenchymal Transition after Exposure to Muscle Invasive Bladder Cancer Exosomes. Oncogenesis 2015, 4, e163. [CrossRef] [PubMed]
  3. Xiao, D.; Barry, S.; Kmetz, D.; Egger, M.; Pan, J.; Rai, S.N.; Qu, J.; McMasters, K.M.; Hao, H. Melanoma Cell-Derived Exosomes Promote Epithelial-Mesenchymal Transition in Primary Melanocytes through Paracrine/Autocrine Signaling in the Tumor Microenvironment. Cancer Lett. 2016, 376, 318–327. [CrossRef]
  4. Santi, A.; Caselli, A.; Ranaldi, F.; Paoli, P.; Mugnaioni, C.; Michelucci, E.; Cirri, P. Cancer Associated Fibroblasts Transfer Lipids and Proteins to Cancer Cells through Cargo Vesicles Supporting Tumor Growth. Biochim. Biophys. Acta 2015, 1853, 3211–3223. [CrossRef]
  5. Luga, V.; Zhang, L.; Viloria-Petit, A.M.; Ogunjimi, A.A.; Inanlou, M.R.; Chiu, E.; Buchanan, M.; Hosein, A.N.; Basik, M.; Wrana, J.L. Exosomes Mediate Stromal Mobilization of Autocrine Wnt-PCP Signaling in Breast Cancer Cell Migration. Cell 2012, 151, 1542–1556. [CrossRef] [PubMed]
  6. Boelens, M.C.; Wu, T.J.; Nabet, B.Y.; Xu, B.; Qiu, Y.; Yoon, T.; Azzam, D.J.; Twyman-Saint Victor, C.; Wiemann, B.Z.; Ishwaran, H.; et al. Exosome Transfer from Stromal to Breast Cancer Cells Regulates Therapy Resistance Pathways. Cell 2014, 159, 499–513. [CrossRef] [PubMed]
  7. Richards, K.E.; Zeleniak, A.E.; Fishel, M.L.; Wu, J.; Littlepage, L.E.; Hill, R. Cancer-Associated Fibroblast Exosomes Regulate Survival and Proliferation of Pancreatic Cancer Cells. Oncogene 2017, 36, 1770–1778. [CrossRef] [PubMed]
  8. Au Yeung, C.L.; Co, N.-N.; Tsuruga, T.; Yeung, T.-L.; Kwan, S.-Y.; Leung, C.S.; Li, Y.; Lu, E.S.; Kwan, K.; Wong, K.-K.; et al. Exosomal Transfer of Stroma-Derived MiR21 Confers Paclitaxel Resistance in Ovarian Cancer Cells through Targeting APAF1. Nat. Commun. 2016, 7, 11150. [CrossRef] [PubMed]
  9. Cui, H.; Seubert, B.; Stahl, E.; Dietz, H.; Reuning, U.; Moreno-Leon, L.; Ilie, M.; Hofman, P.; Nagase, H.; Mari, B.; et al. Tissue Inhibitor of Metalloproteinases-1 Induces a pro-Tumourigenic Increase of MiR-210 in Lung Adenocarcinoma Cells and Their Exosomes. Oncogene 2015, 34, 3640–3650. [CrossRef] [PubMed]
  10. Dang, K.; Myers, K.A. The Role of Hypoxia-Induced MiR-210 in Cancer Progression. Int. J. Mol. Sci. 2015, 16, 6353–6372. [CrossRef] [PubMed]
  11. Mao, G.; Liu, Y.; Fang, X.; Liu, Y.; Fang, L.; Lin, L.; Liu, X.; Wang, N. Tumor-Derived MicroRNA-494 Promotes Angiogenesis in Non-Small Cell Lung Cancer. Angiogenesis 2015, 18, 373–382. [CrossRef]
  12. Matsuura, Y.; Wada, H.; Eguchi, H.; Gotoh, K.; Kobayashi, S.; Kinoshita, M.; Kubo, M.; Hayashi, K.; Iwagami, Y.; Yamada, D.; et al. Exosomal MiR-155 Derived from Hepatocellular Carcinoma Cells Under Hypoxia Promotes Angiogenesis in Endothelial Cells. Dig. Dis. Sci. 2019, 64, 792–802. [CrossRef] [PubMed]
  13. Zhou, X.; Yan, T.; Huang, C.; Xu, Z.; Wang, L.; Jiang, E.; Wang, H.; Chen, Y.; Liu, K.; Shao, Z.; et al. Melanoma Cell-Secreted Exosomal MiR-155-5p Induce Proangiogenic Switch of Cancer-Associated Fibroblasts via SOCS1/JAK2/STAT3 Signaling Pathway. J. Exp. Clin. Cancer Res. 2018, 37, 242. [CrossRef]
  14. Costa-Silva, B.; Aiello, N.M.; Ocean, A.J.; Singh, S.; Zhang, H.; Thakur, B.K.; Becker, A.; Hoshino, A.; Mark, M.T.; Molina, H.; et al. Pancreatic Cancer Exosomes Initiate Pre-Metastatic Niche Formation in the Liver. Nat. Cell Biol. 2015, 17, 816–826. [CrossRef]
  15. Grange, C.; Tapparo, M.; Collino, F.; Vitillo, L.; Damasco, C.; Deregibus, M.C.; Tetta, C.; Bussolati, B.; Camussi, G. Microvesicles Released from Human Renal Cancer Stem Cells Stimulate Angiogenesis and Formation of Lung Premetastatic Niche. Cancer Res. 2011, 71, 5346–5356. [CrossRef] [PubMed]
  16. Hoshino, A.; Costa-Silva, B.; Shen, T.-L.; Rodrigues, G.; Hashimoto, A.; Tesic Mark, M.; Molina, H.; Kohsaka, S.; Di Giannatale, A.; Ceder, S.; et al. Tumour Exosome Integrins Determine Organotropic Metastasis. Nature 2015, 527, 329–335. [CrossRef] [PubMed]
  17. Peinado, H.; Alecˇkovic ́, M.; Lavotshkin, S.; Matei, I.; Costa-Silva, B.; Moreno-Bueno, G.; Hergueta-Redondo, M.; Williams, C.; García-Santos, G.; Ghajar, C.; et al. Melanoma Exosomes Educate Bone Marrow Progenitor Cells toward a Pro-Metastatic Phenotype through MET. Nat. Med. 2012, 18, 883–891. [CrossRef] [PubMed]
  18. Abd Elmageed, Z.Y.; Yang, Y.; Thomas, R.; Ranjan, M.; Mondal, D.; Moroz, K.; Fang, Z.; Rezk, B.M.; Moparty, K.; Sikka, S.C.; et al. Neoplastic Reprogramming of Patient-Derived Adipose Stem Cells by Prostate Cancer Cell-Associated Exosomes. Stem Cells 2014, 32, 983–997. [CrossRef] [PubMed]

Cancers 2019, 11, 891 16 of 21

  1. 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–E2116. [CrossRef] [PubMed]
  2. Chow, A.; Zhou, W.; Liu, L.; Fong, M.Y.; Champer, J.; Van Haute, D.; Chin, A.R.; Ren, X.; Gugiu, B.G.; Meng, Z.; et al. Macrophage Immunomodulation by Breast Cancer-Derived Exosomes Requires Toll-like Receptor 2-Mediated Activation of NF-KB. Sci. Rep. 2014, 4, 5750. [CrossRef] [PubMed]
  3. Marton, A.; Vizler, C.; Kusz, E.; Temesfoi, V.; Szathmary, Z.; Nagy, K.; Szegletes, Z.; Varo, G.; Siklos, L.; Katona, R.L.; et al. Melanoma Cell-Derived Exosomes Alter Macrophage and Dendritic Cell Functions in Vitro. Immunol. Lett. 2012, 148, 34–38. [CrossRef]
  4. Grange, C.; Tapparo, M.; Tritta, S.; Deregibus, M.C.; Battaglia, A.; Gontero, P.; Frea, B.; Camussi, G. Role of HLA-G and Extracellular Vesicles in Renal Cancer Stem Cell-Induced Inhibition of Dendritic Cell Differentiation. BMC Cancer 2015, 15, 1009. [CrossRef]
  5. 101 Ding, G.; Zhou, L.; Qian, Y.; Fu, M.; Chen, J.; Chen, J.; Xiang, J.; Wu, Z.; Jiang, G.; Cao, L. Pancreatic Cancer-Derived Exosomes Transfer MiRNAs to Dendritic Cells and Inhibit RFXAP Expression via MiR-212-3p. Oncotarget 2015, 6, 29877–29888. [CrossRef]
  6. 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. [CrossRef]
  7. Valenti, R.; Huber, V.; Filipazzi, P.; Pilla, L.; Sovena, G.; Villa, A.; Corbelli, A.; Fais, S.; Parmiani, G.; Rivoltini, L. Human Tumor-Released Microvesicles Promote the Differentiation of Myeloid Cells with Transforming Growth Factor-Beta-Mediated Suppressive Activity on T Lymphocytes. Cancer Res. 2006, 66, 9290–9298. [CrossRef] [PubMed]
  8. Chen, G.; Huang, A.C.; Zhang, W.; Zhang, G.; Wu, M.; Xu, W.; Yu, Z.; Yang, J.; Wang, B.; Sun, H.; et al. Exosomal PD-L1 Contributes to Immunosuppression and Is Associated with Anti-PD-1 Response. Nature 2018, 560, 382–386. [CrossRef] [PubMed]
  9. Theodoraki, M.-N.; Yerneni, S.S.; Hoffmann, T.K.; Gooding, W.E.; Whiteside, T.L. Clinical Significance of PD-L1+ Exosomes in Plasma of Head and Neck Cancer Patients. Clin. Cancer Res. 2018, 24, 896–905. [CrossRef] [PubMed]
  10. Yen, E.-Y.; Miaw, S.-C.; Yu, J.-S.; Lai, I.-R. Exosomal TGF-B1 Is Correlated with Lymphatic Metastasis of Gastric Cancers. Am. J. Cancer Res. 2017, 7, 2199–2208. [PubMed]
  11. Abusamra, A.J.; Zhong, Z.; Zheng, X.; Li, M.; Ichim, T.E.; Chin, J.L.; Min, W.-P. Tumor Exosomes Expressing Fas Ligand Mediate CD8+ T-Cell Apoptosis. Blood Cells Mol. Dis. 2005, 35, 169–173. [CrossRef] [PubMed]
  12. Huber, V.; Fais, S.; Iero, M.; Lugini, L.; Canese, P.; Squarcina, P.; Zaccheddu, A.; Colone, M.; Arancia, G.; Gentile, M.; et al. Human Colorectal Cancer Cells Induce T-Cell Death through Release of Proapoptotic Microvesicles: Role in Immune Escape. Gastroenterology 2005, 128, 1796–1804. [CrossRef] [PubMed]
  13. Ashiru, O.; Boutet, P.; Fernández-Messina, L.; Agüera-González, S.; Skepper, J.N.; Valés-Gómez, M.; Reyburn, H.T. Natural Killer Cell Cytotoxicity Is Suppressed by Exposure to the Human NKG2D Ligand MICA*008 That Is Shed by Tumor Cells in Exosomes. Cancer Res. 2010, 70, 481–489. [CrossRef] [PubMed]
  14. 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. [CrossRef]
  15. Ludwig, S.; Floros, T.; Theodoraki, M.-N.; Hong, C.-S.; Jackson, E.K.; Lang, S.; Whiteside, T.L. Suppression of Lymphocyte Functions by Plasma Exosomes Correlates with Disease Activity in Patients with Head and Neck Cancer. Clin. Cancer Res. 2017, 23, 4843–4854. [CrossRef]
  16. Escudier,B.;Dorval,T.;Chaput,N.;André,F.;Caby,M.-P.;Novault,S.;Flament,C.;Leboulaire,C.;Borg,C.; Amigorena, S.; et al. Vaccination of Metastatic Melanoma Patients with Autologous Dendritic Cell (DC) Derived-Exosomes: Results of Thefirst Phase I Clinical Trial. J. Transl. Med. 2005, 3, 10. [CrossRef]
  17. Morse,M.A.;Garst,J.;Osada,T.;Khan,S.;Hobeika,A.;Clay,T.M.;Valente,N.;Shreeniwas,R.;Sutton,M.A.; Delcayre, A.; et al. A Phase I Study of Dexosome Immunotherapy in Patients with Advanced Non-Small Cell Lung Cancer. J. Transl. Med. 2005, 3, 9. [CrossRef] [PubMed]
  18. Taylor,D.D.;Gercel-Taylor,C.MicroRNASignaturesofTumor-DerivedExosomesasDiagnosticBiomarkers of Ovarian Cancer. Gynecol. Oncol. 2008, 110, 13–21. [CrossRef] [PubMed]

Cancers 2019, 11, 891 17 of 21

  1. Rabinowits,G.;Gerçel-Taylor,C.;Day,J.M.;Taylor,D.D.;Kloecker,G.H.ExosomalMicroRNA:ADiagnostic Marker for Lung Cancer. Clin. Lung Cancer 2009, 10, 42–46. [CrossRef] [PubMed]
  2. Sohn, W.; Kim, J.; Kang, S.H.; Yang, S.R.; Cho, J.-Y.; Cho, H.C.; Shim, S.G.; Paik, Y.-H. Serum Exosomal MicroRNAs as Novel Biomarkers for Hepatocellular Carcinoma. Exp. Mol. Med. 2015, 47, e184. [CrossRef] [PubMed]
  3. Taylor,D.D.;Gercel-Taylor,C.Exosomes/Microvesicles:MediatorsofCancer-AssociatedImmunosuppressive Microenvironments. Semin. Immunopathol. 2011, 33, 441–454. [CrossRef]
  4. Jabalee,J.;Towle,R.;Garnis,C.TheRoleofExtracellularVesiclesinCancer:Cargo,Function,andTherapeutic Implications. Cells 2018, 7, 93. [CrossRef] [PubMed]
  5. Bu,N.;Wu,H.;Sun,B.;Zhang,G.;Zhan,S.;Zhang,R.;Zhou,L.Exosome-LoadedDendriticCellsElicit Tumor-Specific CD8+ Cytotoxic T Cells in Patients with Glioma. J. Neurooncol. 2011, 104, 659–667. [CrossRef]
  6. Graner, M.W.; Alzate, O.; Dechkovskaia, A.M.; Keene, J.D.; Sampson, J.H.; Mitchell, D.A.; Bigner, D.D. Proteomic and Immunologic Analyses of Brain Tumor Exosomes. FASEB J. 2009, 23, 1541–1557. [CrossRef]
  7. Kunigelis,K.E.;Graner,M.W.TheDichotomyofTumorExosomes(TEX)inCancerImmunity:IsItAllinthe
    ConTEXt? Vaccines (Basel) 2015, 3, 1019–1051. [CrossRef]
  8. Chevillet, J.R.; Kang, Q.; Ruf, I.K.; Briggs, H.A.; Vojtech, L.N.; Hughes, S.M.; Cheng, H.H.; Arroyo, J.D.;
    Meredith, E.K.; Gallichotte, E.N.; et al. Quantitative and Stoichiometric Analysis of the MicroRNA Content
    of Exosomes. Proc. Natl. Acad. Sci. USA 2014, 111, 14888–14893. [CrossRef]
  9. Anfossi, S.; Giordano, A.; Gao, H.; Cohen, E.N.; Tin, S.; Wu, Q.; Garza, R.J.; Debeb, B.G.; Alvarez, R.H.;
    Valero, V.; et al. High Serum MiR-19a Levels Are Associated with Inflammatory Breast Cancer and Are Predictive of Favorable Clinical Outcome in Patients with Metastatic HER2+ Inflammatory Breast Cancer. PLoS ONE 2014, 9, e83113. [CrossRef] [PubMed]
  10. Gemmell, C.H.; Sefton, M.V.; Yeo, E.L. Platelet-Derived Microparticle Formation Involves Glycoprotein IIb-IIIa. Inhibition by RGDS and a Glanzmann’s Thrombasthenia Defect. J. Biol. Chem. 1993, 268, 14586–14589. [PubMed]
  11. Yan, W.; Apweiler, R.; Balgley, B.M.; Boontheung, P.; Bundy, J.L.; Cargile, B.J.; Cole, S.; Fang, X.; Gonzalez-Begne, M.; Griffin, T.J.; et al. Systematic Comparison of the Human Saliva and Plasma Proteomes. Proteom. Clin. Appl. 2009, 3, 116–134. [CrossRef] [PubMed]
  12. Bandhakavi,S.;Stone,M.D.;Onsongo,G.;VanRiper,S.K.;Griffin,T.J.ADynamicRangeCompressionand Three-Dimensional Peptide Fractionation Analysis Platform Expands Proteome Coverage and the Diagnostic Potential of Whole Saliva. J. Proteome Res. 2009, 8, 5590–5600. [CrossRef]
  13. Zhao, M.; Yang, Y.; Guo, Z.; Shao, C.; Sun, H.; Zhang, Y.; Sun, Y.; Liu, Y.; Song, Y.; Zhang, L.; et al. A Comparative Proteomics Analysis of Five Body Fluids: Plasma, Urine, Cerebrospinal Fluid, Amniotic Fluid, and Saliva. Proteom. Clin. Appl. 2018, 12, e1800008. [CrossRef]
  14. Omenn,G.S.;States,D.J.;Adamski,M.;Blackwell,T.W.;Menon,R.;Hermjakob,H.;Apweiler,R.;Haab,B.B.; Simpson, R.J.; Eddes, J.S.; et al. Overview of the HUPO Plasma Proteome Project: Results from the Pilot Phase with 35 Collaborating Laboratories and Multiple Analytical Groups, Generating a Core Dataset of 3020 Proteins and a Publicly-Available Database. Proteomics 2005, 5, 3226–3245. [CrossRef]
  15. Cheng,Y.;Pereira,M.;Raukar,N.;Reagan,J.L.;Queseneberry,M.;Goldberg,L.;Borgovan,T.;LaFrance,W.C.; Dooner, M.; Deregibus, M.; et al. Potential Biomarkers to Detect Traumatic Brain Injury by the Profiling of Salivary Extracellular Vesicles. J. Cell. Physiol. 2019, 234, 14377–14388. [CrossRef]
  16. Saeedi, S.; Israel, S.; Nagy, C.; Turecki, G. The Emerging Role of Exosomes in Mental Disorders. Transl. Psychiatry 2019, 9, 122. [CrossRef]
  17. Cao,Z.;Wu,Y.;Liu,G.;Jiang,Y.;Wang,X.;Wang,Z.;Feng,T.α-SynucleininSalivaryExtracellularVesicles as a Potential Biomarker of Parkinson’s Disease. Neurosci. Lett. 2019, 696, 114–120. [CrossRef]
  18. Jin,Y.;Guan,Z.;Wang,X.;Wang,Z.;Zeng,R.;Xu,L.;Cao,P.ALA-PDTPromotesHPV-PositiveCervical Cancer Cells Apoptosis and DCs Maturation via MiR-34a Regulated HMGB1 Exosomes Secretion. Photodiagn. Photodyn. 2018, 24, 27–35. [CrossRef]
  19. Harden,M.E.;Munger,K.HumanPapillomavirus16E6andE7OncoproteinExpressionAltersMicroRNA Expression in Extracellular Vesicles. Virology 2017, 508, 63–69. [CrossRef] [PubMed]
  20. Honegger,A.;Leitz,J.;Bulkescher,J.;Hoppe-Seyler,K.;Hoppe-Seyler,F.SilencingofHumanPapillomavirus (HPV) E6/E7 Oncogene Expression Affects Both the Contents and the Amounts of Extracellular Microvesicles Released from HPV-Positive Cancer Cells. Int. J. Cancer 2013, 133, 1631–1642. [CrossRef] [PubMed]

Cancers 2019, 11, 891 18 of 21

  1. Chiantore,M.V.;Mangino,G.;Iuliano,M.;Zangrillo,M.S.;DeLillis,I.;Vaccari,G.;Accardi,R.;Tommasino,M.; Columba Cabezas, S.; Federico, M.; et al. Human Papillomavirus E6 and E7 Oncoproteins Affect the Expression of Cancer-Related MicroRNAs: Additional Evidence in HPV-Induced Tumorigenesis. J. Cancer Res. Clin. Oncol. 2016, 142, 1751–1763. [CrossRef] [PubMed]
  2. Peacock,B.;Rigby,A.;Bradford,J.;Pink,R.;Hunter,K.;Lambert,D.;Hunt,S.ExtracellularVesicleMicroRNA Cargo Is Correlated with HPV Status in Oropharyngeal Carcinoma. J. Oral Pathol. Med. 2018, 47, 954–963. [CrossRef] [PubMed]
  3. Hukin,J.;Farrell,K.;MacWilliam,L.M.;Colbourne,M.;Waida,E.;Tan,R.;Mroz,L.;Thomas,E.Case-Control Study of Primary Human Herpesvirus 6 Infection in Children with Febrile Seizures. Pediatrics 1998, 101, E3. [CrossRef] [PubMed]
  4. Sharma,S.;Gillespie,B.M.;Palanisamy,V.;Gimzewski,J.K.QuantitativeNanostructuralandSingle-Molecule Force Spectroscopy Biomolecular Analysis of Human-Saliva-Derived Exosomes. Langmuir 2011, 27, 14394–14400. [CrossRef] [PubMed]
  5. Zlotogorski-Hurvitz,A.;Dayan,D.;Chaushu,G.;Salo,T.;Vered,M.MorphologicalandMolecularFeatures of Oral Fluid-Derived Exosomes: Oral Cancer Patients versus Healthy Individuals. J. Cancer Res. Clin. Oncol. 2016, 142, 101–110. [CrossRef] [PubMed]
  6. Winck, F.V.; Prado Ribeiro, A.C.; Ramos Domingues, R.; Ling, L.Y.; Riaño-Pachón, D.M.; Rivera, C.; Brandão, T.B.; Gouvea, A.F.; Santos-Silva, A.R.; Coletta, R.D.; et al. Insights into Immune Responses in Oral Cancer through Proteomic Analysis of Saliva and Salivary Extracellular Vesicles. Sci. Rep. 2015, 5, 16305. [CrossRef] [PubMed]
  7. Gai, C.; Camussi, F.; Broccoletti, R.; Gambino, A.; Cabras, M.; Molinaro, L.; Carossa, S.; Camussi, G.; Arduino, P.G. Salivary Extracellular Vesicle-Associated MiRNAs as Potential Biomarkers in Oral Squamous Cell Carcinoma. BMC Cancer 2018, 18, 439. [CrossRef] [PubMed]
  8. Sun,Y.;Xia,Z.;Shang,Z.;Sun,K.;Niu,X.;Qian,L.;Fan,L.-Y.;Cao,C.-X.;Xiao,H.FacilePreparationof Salivary Extracellular Vesicles for Cancer Proteom. Sci. Rep. 2016, 6, 24669. [CrossRef] [PubMed]
  9. Sun, Y.; Huo, C.; Qiao, Z.; Shang, Z.; Uzzaman, A.; Liu, S.; Jiang, X.; Fan, L.-Y.; Ji, L.; Guan, X.; et al. Comparative Proteomic Analysis of Exosomes and Microvesicles in Human Saliva for Lung Cancer. J. Proteome Res. 2018, 17, 1101–1107. [CrossRef] [PubMed]
  10. Langevin, S.; Kuhnell, D.; Parry, T.; Biesiada, J.; Huang, S.; Wise-Draper, T.; Casper, K.; Zhang, X.; Medvedovic, M.; Kasper, S. Comprehensive MicroRNA-Sequencing of Exosomes Derived from Head and Neck Carcinoma Cells in Vitro Reveals Common Secretion Profiles and Potential Utility as Salivary Biomarkers. Oncotarget 2017, 8, 82459–82474. [CrossRef] [PubMed]
  11. Machida,T.;Tomofuji,T.;Maruyama,T.;Yoneda,T.;Ekuni,D.;Azuma,T.;Miyai,H.;Mizuno,H.;Kato,H.; Tsutsumi, K.; et al. MiR-1246 and MiR-4644 in Salivary Exosome as Potential Biomarkers for Pancreatobiliary Tract Cancer. Oncol. Rep. 2016, 36, 2375–2381. [CrossRef] [PubMed]
  12. Lau,C.;Kim,Y.;Chia,D.;Spielmann,N.;Eibl,G.;Elashoff,D.;Wei,F.;Lin,Y.-L.;Moro,A.;Grogan,T.;etal. Role of Pancreatic Cancer-Derived Exosomes in Salivary Biomarker Development. J. Biol. Chem. 2013, 288, 26888–26897. [CrossRef] [PubMed]
  13. Kowal,J.;Arras,G.;Colombo,M.;Jouve,M.;Morath,J.P.;Primdal-Bengtson,B.;Dingli,F.;Loew,D.;Tkach,M.; Théry, C. Proteomic Comparison Defines Novel Markers to Characterize Heterogeneous Populations of Extracellular Vesicle Subtypes. Proc. Natl. Acad. Sci. USA 2016, 113, E968–E977. [CrossRef] [PubMed]
  14. Momen-Heravi, F.; Balaj, L.; Alian, S.; Trachtenberg, A.J.; Hochberg, F.H.; Skog, J.; Kuo, W.P. Impact of Biofluid Viscosity on Size and Sedimentation Efficiency of the Isolated Microvesicles. Front. Physiol. 2012, 3, 162. [CrossRef] [PubMed]
  15. Jeppesen,D.K.;Hvam,M.L.;Primdahl-Bengtson,B.;Boysen,A.T.;Whitehead,B.;Dyrskjøt,L.;Orntoft,T.F.; Howard, K.A.; Ostenfeld, M.S. Comparative Analysis of Discrete Exosome Fractions Obtained by Differential Centrifugation. J. Extracell. Vesicles 2014, 3, 25011. [CrossRef] [PubMed]
  16. Cvjetkovic,A.;Lötvall,J.;Lässer,C.TheInfluenceofRotorTypeandCentrifugationTimeontheYieldand Purity of Extracellular Vesicles. J. Extracell. Vesicles 2014, 3, 23111. [CrossRef] [PubMed]
  17. Tauro,B.J.;Greening,D.W.;Mathias,R.A.;Ji,H.;Mathivanan,S.;Scott,A.M.;Simpson,R.J.Comparison of Ultracentrifugation, Density Gradient Separation, and Immunoaffinity Capture Methods for Isolating Human Colon Cancer Cell Line LIM1863-Derived Exosomes. Methods 2012, 56, 293–304. [CrossRef]

Cancers 2019, 11, 891 19 of 21

  1. Müller,G.NovelToolsfortheStudyofCellType-SpecificExosomesandMicrovesicles.J.Bioanal.Biomed. 2012, 4, 46–60. [CrossRef]
  2. Taylor,D.D.;Lyons,K.S.;Gerçel-Taylor,C.ShedMembraneFragment-AssociatedMarkersforEndometrial and Ovarian Cancers. Gynecol. Oncol. 2002, 84, 443–448. [CrossRef] [PubMed]
  3. Böing, A.N.; van der Pol, E.; Grootemaat, A.E.; Coumans, F.A.W.; Sturk, A.; Nieuwland, R. Single-Step Isolation of Extracellular Vesicles by Size-Exclusion Chromatography. J. Extracell. Vesicles 2014, 3, 23430. [CrossRef] [PubMed]
  4. Taylor,D.D.;Shah,S.MethodsofIsolatingExtracellularVesiclesImpactDown-StreamAnalysesofTheir Cargoes. Methods 2015, 87, 3–10. [CrossRef] [PubMed]
  5. Caby,M.-P.;Lankar,D.;Vincendeau-Scherrer,C.;Raposo,G.;Bonnerot,C.Exosomal-likeVesiclesArePresent in Human Blood Plasma. Int. Immunol. 2005, 17, 879–887. [CrossRef] [PubMed]
  6. Zarovni,N.;Corrado,A.;Guazzi,P.;Zocco,D.;Lari,E.;Radano,G.;Muhhina,J.;Fondelli,C.;Gavrilova,J.; Chiesi, A. Integrated Isolation and Quantitative Analysis of Exosome Shuttled Proteins and Nucleic Acids Using Immunocapture Approaches. Methods 2015, 87, 46–58. [CrossRef] [PubMed]
  7. Szatanek,R.;Baran,J.;Siedlar,M.;Baj-Krzyworzeka,M.IsolationofExtracellularVesicles:Determiningthe Correct Approach (Review). Int. J. Mol. Med. 2015, 36, 11–17. [CrossRef] [PubMed]
  8. Alvarez,M.L.;Khosroheidari,M.;KanchiRavi,R.;DiStefano,J.K.ComparisonofProtein,MicroRNA,and MRNA Yields Using Different Methods of Urinary Exosome Isolation for the Discovery of Kidney Disease Biomarkers. Kidney Int. 2012, 82, 1024–1032. [CrossRef]
  9. Alvarez,M.L.IsolationofUrinaryExosomesforRNABiomarkerDiscoveryUsingaSimple,Fast,andHighly Scalable Method. Methods Mol. Biol. 2014, 1182, 145–170. [CrossRef]
  10. Rekker,K.;Saare,M.;Roost,A.M.;Kubo,A.-L.;Zarovni,N.;Chiesi,A.;Salumets,A.;Peters,M.Comparison of Serum Exosome Isolation Methods for MicroRNA Profiling. Clin. Biochem. 2014, 47, 135–138. [CrossRef]
  11. Zlotogorski-Hurvitz, A.; Dayan, D.; Chaushu, G.; Korvala, J.; Salo, T.; Sormunen, R.; Vered, M. Human
    Saliva-Derived Exosomes: Comparing Methods of Isolation. J. Histochem. Cytochem. 2015, 63, 181–189.
    [CrossRef]
  12. Kanchi Ravi, R.; Khosroheidari, M.; DiStefano, J.K. A Modified Precipitation Method to Isolate Urinary
    Exosomes. J. Vis. Exp. 2015, 95, e51158. [CrossRef] [PubMed]
  13. Witwer, K.W.; Buzás, E.I.; Bemis, L.T.; Bora, A.; Lässer, C.; Lötvall, J.; Nolte-’t Hoen, E.N.; Piper, M.G.;
    Sivaraman, S.; Skog, J.; et al. Standardization of Sample Collection, Isolation and Analysis Methods in
    Extracellular Vesicle Research. J. Extracell. Vesicles 2013, 2, 20360. [CrossRef]
  14. Yuana,Y.;Levels,J.;Grootemaat,A.;Sturk,A.;Nieuwland,R.Co-IsolationofExtracellularVesiclesand
    High-Density Lipoproteins Using Density Gradient Ultracentrifugation. J. Extracell. Vesicles 2014, 3, 23262.
    [CrossRef] [PubMed]
  15. Wei,F.;Lin,C.-C.;Joon,A.;Feng,Z.;Troche,G.;Lira,M.E.;Chia,D.;Mao,M.;Ho,C.-L.;Su,W.-C.;etal.
    Noninvasive Saliva-Based EGFR Gene Mutation Detection in Patients with Lung Cancer. Am. J. Respir. Crit.
    Care Med. 2014, 190, 1117–1126. [CrossRef] [PubMed]
  16. Pu,D.;Liang,H.;Wei,F.;Akin,D.;Feng,Z.;Yan,Q.;Li,Y.;Zhen,Y.;Xu,L.;Dong,G.;etal.Evaluationofa
    Novel Saliva-Based Epidermal Growth Factor Receptor Mutation Detection for Lung Cancer: A Pilot Study.
    Thorac. Cancer 2016, 7, 428–436. [CrossRef]
  17. Wei, F.; Yang, J.; Wong, D.T.W. Detection of Exosomal Biomarker by Electric Field-Induced Release and
    Measurement (EFIRM). Biosens. Bioelectron. 2013, 44, 115–121. [CrossRef] [PubMed]
  18. Cheng, J.; Nonaka, T.; Wong, D.T.W. Salivary Exosomes as Nanocarriers for Cancer Biomarker Delivery.
    Materials (Basel) 2019, 12, 654. [CrossRef]
  19. Michael, A.; Bajracharya, S.D.; Yuen, P.S.T.; Zhou, H.; Star, R.A.; Illei, G.G.; Alevizos, I. Exosomes from
    Human Saliva as a Source of MicroRNA Biomarkers. Oral Dis. 2010, 16, 34–38. [CrossRef]
  20. Deregibus,M.C.;Figliolini,F.;D’Antico,S.;Manzini,P.M.;Pasquino,C.;DeLena,M.;Tetta,C.;Brizzi,M.F.; Camussi, G. Charge-Based Precipitation of Extracellular Vesicles. Int. J. Mol. Med. 2016, 38, 1359–1366.
    [CrossRef]
  21. Li,Y.;StJohn,M.A.R.;Zhou,X.;Kim,Y.;Sinha,U.;Jordan,R.C.K.;Eisele,D.;Abemayor,E.;Elashoff,D.;
    Park, N.-H.; et al. Salivary Transcriptome Diagnostics for Oral Cancer Detection. Clin. Cancer Res. 2004, 10, 8442–8450. [CrossRef]

Cancers 2019, 11, 891 20 of 21

  1. Zhou, Y.; Kolokythas, A.; Schwartz, J.L.; Epstein, J.B.; Adami, G.R. MicroRNA from Brush Biopsy to Characterize Oral Squamous Cell Carcinoma Epithelium. Cancer Med. 2017, 6, 67–78. [CrossRef]
  2. Yap,T.;Vella,L.J.;Seers,C.;Nastri,A.;Reynolds,E.;Cirillo,N.;McCullough,M.OralSwirlSamples—A Robust Source of MicroRNA Protected by Extracellular Vesicles. Oral Dis. 2017, 23, 312–317. [CrossRef]
  3. Mestdagh,P.;VanVlierberghe,P.;DeWeer,A.;Muth,D.;Westermann,F.;Speleman,F.;Vandesompele,J. A Novel and Universal Method for MicroRNA RT-QPCR Data Normalization. Genome Biol. 2009, 10, R64. [CrossRef]
  4. Kaczor-Urbanowicz,K.E.;MartinCarreras-Presas,C.;Aro,K.;Tu,M.;Garcia-Godoy,F.;Wong,D.T.Saliva Diagnostics-Current Views and Directions. Exp. Biol. Med. (Maywood) 2017, 242, 459–472. [CrossRef]
  5. Zhang, L.; Farrell, J.J.; Zhou, H.; Elashoff, D.; Akin, D.; Park, N.-H.; Chia, D.; Wong, D.T. Salivary
    Transcriptomic Biomarkers for Detection of Resectable Pancreatic Cancer. Gastroenterology 2010, 138, e1–e7.
    [CrossRef]
  6. Lee,Y.-H.;Kim,J.H.;Zhou,H.;Kim,B.W.;Wong,D.T.SalivaryTranscriptomicBiomarkersforDetectionof
    Ovarian Cancer: For Serous Papillary Adenocarcinoma. J. Mol. Med. 2012, 90, 427–434. [CrossRef]
  7. Xiao,H.;Zhang,L.;Zhou,H.;Lee,J.M.;Garon,E.B.;Wong,D.T.W.ProteomicAnalysisofHumanSaliva from Lung Cancer Patients Using Two-Dimensional Difference Gel Electrophoresis and Mass Spectrometry.
    Mol. Cell Proteom. 2012, 11, M111.012112. [CrossRef]
  8. Ogawa, Y.; Tsujimoto, M.; Yanoshita, R. Next-Generation Sequencing of Protein-Coding and Long
    Non-Protein-Coding RNAs in Two Types of Exosomes Derived from Human Whole Saliva. Biol. Pharm. Bull.
    2016, 39, 1496–1507. [CrossRef]
  9. Palanisamy, V.; Sharma, S.; Deshpande, A.; Zhou, H.; Gimzewski, J.; Wong, D.T. Nanostructural and
    Transcriptomic Analyses of Human Saliva Derived Exosomes. PLoS ONE 2010, 5, e8577. [CrossRef]
  10. Lau,C.S.;Wong,D.T.W.BreastCancerExosome-likeMicrovesiclesandSalivaryGlandCellsInterplayAlters
    Salivary Gland Cell-Derived Exosome-like Microvesicles in Vitro. PLoS ONE 2012, 7, e33037. [CrossRef]
  11. Katsiougiannis, S.; Chia, D.; Kim, Y.; Singh, R.P.; Wong, D.T.W. Saliva Exosomes from Pancreatic Tumor-Bearing Mice Modulate NK Cell Phenotype and Antitumor Cytotoxicity. FASEB J. 2017, 31, 998–1010.
    [CrossRef]
  12. Serban,K.A.;Rezania,S.;Petrusca,D.N.;Poirier,C.;Cao,D.;Justice,M.J.;Patel,M.;Tsvetkova,I.;Kamocki,K.;
    Mikosz, A.; et al. Structural and Functional Characterization of Endothelial Microparticles Released by
    Cigarette Smoke. Sci. Rep. 2016, 6, 31596. [CrossRef]
  13. Kodidela,S.;Wang,Y.;Patters,B.J.;Gong,Y.;Sinha,N.;Ranjit,S.;Gerth,K.;Haque,S.;Cory,T.;McArthur,C.;
    et al. Proteomic Profiling of Exosomes Derived from Plasma of HIV-Infected Alcohol Drinkers and Cigarette
    Smokers. J. Neuroimmune Pharmacol. 2019, 1–19. [CrossRef]
  14. Appert-Collin,A.;Hubert,P.;Crémel,G.;Bennasroune,A.RoleofErbBReceptorsinCancerCellMigration
    and Invasion. Front. Pharm. 2015, 6, 283. [CrossRef]
  15. Li,X.;Sun,R.;Geng,X.;Wang,S.;Zen,D.;Pei,J.;Yang,J.;Fan,Y.;Jiang,H.;Yang,P.;etal.AComprehensive
    Analysis of Candidate Gene Signatures in Oral Squamous Cell Carcinoma. Neoplasma 2017, 64, 167–174.
    [CrossRef]
  16. Ohnishi,Y.;Yasui,H.;Kakudo,K.;Nozaki,M.Lapatinib-ResistantCancerCellsPossessingEpithelialCancer
    Stem Cell Properties Develop Sensitivity during Sphere Formation by Activation of the ErbB/AKT/Cyclin D2
    Pathway. Oncol. Rep. 2016, 36, 3058–3064. [CrossRef]
  17. Li, S.-X.; Yang, Y.-Q.; Jin, L.-J.; Cai, Z.-G.; Sun, Z. Detection of Survivin, Carcinoembryonic Antigen and
    ErbB2 Level in Oral Squamous Cell Carcinoma Patients. Cancer Biomark. 2016, 17, 377–382. [CrossRef]
  18. Dong,C.;Ye,D.-X.;Zhang,W.-B.;Pan,H.-Y.;Zhang,Z.-Y.;Zhang,L.OverexpressionofC-FosPromotesCell Invasion and Migration via CD44 Pathway in Oral Squamous Cell Carcinoma. J. Oral Pathol. Med. 2015, 44,
    353–360. [CrossRef]
  19. Judd, N.P.; Winkler, A.E.; Murillo-Sauca, O.; Brotman, J.J.; Law, J.H.; Lewis, J.S.; Dunn, G.P.; Bui, J.D.;
    Sunwoo, J.B.; Uppaluri, R. ERK1/2 Regulation of CD44 Modulates Oral Cancer Aggressiveness. Cancer Res.
    2012, 72, 365–374. [CrossRef]
  20. Ghuwalewala, S.; Ghatak, D.; Das, P.; Dey, S.; Sarkar, S.; Alam, N.; Panda, C.K.; Roychoudhury, S.
    CD44(High)CD24(Low) Molecular Signature Determines the Cancer Stem Cell and EMT Phenotype in Oral Squamous Cell Carcinoma. Stem Cell Res. 2016, 16, 405–417. [CrossRef]

Cancers 2019, 11, 891 21 of 21

  1. Meng, W.; Xia, Q.; Wu, L.; Chen, S.; He, X.; Zhang, L.; Gao, Q.; Zhou, H. Downregulation of TGF-Beta Receptor Types II and III in Oral Squamous Cell Carcinoma and Oral Carcinoma-Associated Fibroblasts. BMC Cancer 2011, 11, 88. [CrossRef]
  2. Momen-Heravi,F.;Bala,S.ExtracellularvesiclesinoralsquamouscarcinomacarryoncogenicmiRNAprofile and reprogram monocytes via NF-κB pathway. Oncotarget 2018, 9, 34838–34854. [CrossRef]
  3. Arantes, L.M.R.B.; De Carvalho, A.C.; Melendez, M.E.; Lopes Carvalho, A. Serum, Plasma and Saliva Biomarkers for Head and Neck Cancer. Expert Rev. Mol. Diagn. 2018, 18, 85–112. [CrossRef]
  4. Tang,H.;Wu,Z.;Zhang,J.;Su,B.SalivaryLncRNAasaPotentialMarkerforOralSquamousCellCarcinoma Diagnosis. Mol. Med. Rep. 2013, 7, 761–766. [CrossRef]
    © 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).