5

Molecular Mechanisms of Glioma Cell Motility

Angela Armento* Jakob Ehlers* Sonja Schötterl* Ulrike Naumann

Department of Vascular Neurology, Hertie Institute for Clinical Brain Research and Center Neurology, University of Tübingen, Tübingen, Germany

Abstract: Gliomas are the most common intracranial tumors in humans. The most malignant among these tumors is glioblastoma (GBM), with an incidence of 3–5 out of 100,000 persons in Western countries. GBM arises either de novo (primary GBM) or develops from a lower grade glioma (secondary GBM). The prognosis is poor. GBMs are lethal tumors and even optimal surgical resection, followed by chemotherapy and irradiation, results in a median survival of about 12–15 months. One characteristic that is responsible for GBM malignancy, and its worse prognosis, is the highly infiltrative growth of GBM cells into the healthy brain. GBM cell migration and invasion is a very complex process that is regulated by several factors, which include changes in the migrating cell itself as well as the tumor microenvironment. This chapter provides an overview of routes of invasion of glioma cells, the signaling pathways that drive glioma cell motility, and the processes through which glioma cells modulate their surrounding environment.

Keywords: Glioma migration; Invasion; Molecular mechanisms

Author for correspondence: Ulrike Naumann, Department of Vascular Neurology, Hertie Institute for Clinical Brain Research and Center Neurology, University of Tübingen, Otfried-Müller-Str. 27, DE-72076 Tübingen, Germany. E-mail: ulrike.naumann@uni-tuebingen.de

Doi: http://dx.doi.org/10.15586/codon.glioblastoma.2017.ch5

*These authors contributed equally for authorship.

In: Glioblastoma. Steven De Vleeschouwer (Editor), Codon Publications, Brisbane, Australia ISBN: 978-0-9944381-2-6; Doi: http://dx.doi.org/10.15586/codon.glioblastoma.2017

Copyright: The Authors.

Licence: This open access article is licenced under Creative Commons Attribution-NonCommercial 4.0 International (CC BY 4.0). https://creativecommons.org/licenses/by-nc/4.0/

Introduction

Glioblastoma (GBM), the most malignant brain tumor, has a complex biology, and despite decades of research, much is still unknown. GBM separates itself from lower grade gliomas by exhibiting central necrosis and microvascular proliferation. It is characterized by a rapid and highly infiltrative growth. In GBM, extracranial metastases are extremely rare; tumor cell invasion and migration are the main features of GBM spreading (1). The invasive nature of GBM leads to local destruction of healthy tissues, and is the main source of recurrence (2). Even with the best imaging methods available, it is difficult to detect cells that had migrated away from the primary tumor. Glioma cells are able to migrate far away from the original tumor and can even cross into the contralateral hemisphere making complete surgical resection of GBM impossible (3). Invasion of glioma cells into the healthy brain also leads to the escape of these cells from irradiation and chemotherapy. Therefore, understanding the biology of glioma cell motility is of great importance for developing novel therapeutic approaches to treat GBM patients.

Glioma cells mainly use two routes to invade the healthy brain: the perivascular space around blood vessels and axons (4). Whether glioma cells exclusively use one route over the other, or whether other roads are also utilized, is not fully understood. In addition, it is not known how glioma cells decide to choose one pathway over the other for invasion. There are several cellular and environmental requirements that set the stage for a glioma cell to move. For example, migrating cells show changes in energy metabolism that are often induced by hypoxic conditions (5, 6). Cytokines, chemokines, nutrition deprivation, and hypoxia lead to changes in the expression of transcription factors (TFs), and subsequently to altered protein expression (7). In this regard, differential expression of ion channels, neurotransmitters, proteases, chemokines, and cytokines has been described in moving versus resting glioma cells (2). Besides transcriptional changes, the cytoskeleton of the glioma cell has to be rearranged to allow cell movement, cell adhesion has to be reduced, and the tumor cell has to be shrunk to fit into the small perivascular space. Furthermore, the extracellular matrix (ECM) has to be remodeled or destroyed to allow glioma cell invasion (8). Even the interaction of glioma cells with adjacent nonneoplastic cells like astrocytes or endothelial cells is important for glioma cell migration (9, 10). This chapter gives an overview of different processes and mechanisms glioma cells use to migrate and invade, and the signaling cascades that regulate the motility of glioma cells.

Infiltration of Diffuse Glioma

PATTERNS OF GLIOMA CELL INFILTRATION

Glioma cells infiltrate into the healthy brain parenchyma using preexisting structures like blood vessels or myelinated nerve fibers of white matter tracts, both of which present high mechanical rigidity (11, 12). ECM stiffness is a major regulator of cell motility. The movement of cells toward a more rigid ECM area is called mechanotaxis (13). A more rigid ECM, as in the perivascular space, promotes glioma cell migration (14). Stiffness varies with the grade of glioma. It is known that invasive GBM produces stiffness-promoting factors like collagen, fibronectin (FN), and laminin. Furthermore, glioma cells overexpress components of the basal membrane of the cerebral vasculature, for example, tenascin (TN)-C, which is associated with glioma progression (15). Glioma cells are recruited to the perivascular space around blood vessels by chemoattractants like bradykinin, which is produced by endothelial cells (16). Also, overexpression of chemokine receptors on glioma cells has been associated with perivascular invasion (17). Cell movement along white matter tracts, a second known route of glioma cell invasion, is mediated by a variety of proteins called axonal guidance molecules (see the section “Axonal Guidance Molecules”), which act as attracting or repelling factors.

HYPOXIA

The center of GBM is characterized by necrosis, surrounded by an area where tumor cells deal with hypoxia and nutrient starvation. Around the necrotic region, the population of “pseudopalisading” cells become prominent. These glioma cells activate migratory processes in an attempt to escape hypoxia and to reach oxygen-rich areas adjacent to blood vessels (18). Some of the pro-migratory and pro-invasive factors produced or activated in response to hypoxic conditions include: metalloproteases like MMP-9, A Disintegrin, and Metalloproteinase (ADAM)-17 (19, 20); galectins (21); epithelial to mesenchymal transition (EMT) transcriptional regulators like SLUG and SNAIL and the zinc-finger E-Box-binding homeobox proteins ZEB-1 and ZEB-2 (22, 23); and CXCR4 and CXCR7, the latter mediating glioma cell migration toward stromal-derived factor (SDF)-1α/CXCL12 (24, 25).

THE “GO OR GROW” OF TUMOR CELLS

Migration and proliferation of glioma cells are mutually exclusive. This phenomenon, called “Go or Grow,” was first discovered in astrocytoma cells, where proliferation and migration are timely separated (26). The “Go or Grow” is modulated by changes in the microenvironment like hypoxia or nutrient depletion, which prompts a tumor cell to “Go” in order to reach a more favorable environment and re-settle there, or to “Grow” if the environment provides enough oxygen and nutrients. The pentose phosphate pathway (PPP) is mainly used during proliferation, and glycolysis is used as the energy source during migration (5). Other parameters that influence the “Go or Grow” of glioma cells are the cell volume, cytoskeleton dynamics, and the ECM composition (27). Differential activation of TFs has been reported: increased NF-κB activity in migrating cells, and c-myc in proliferating cells (28). Also, changes in miRNAs expression modulate the “Go or Grow”: elevated miR-451 expression is associated with a shorter GBM patient survival and higher proliferation (29), whereas mir-9, being highly expressed in glioma cells, inhibits proliferation but promotes migration (30). Understanding the process of “Go or Grow” in glioma is of central importance since it is known that ionizing irradiation used for the treatment of GBM promotes the “Go” and thereby the invasive phenotype of glioma cells (31, 32).

Extracellular Matrix

ECM constitutes 10–20% of brain volume. It is produced by the surrounding cells. ECM not only has a structural function but also a major role in brain development, cell survival, migration, maturation, differentiation, and tissue homeostasis (33, 34). The main components of the brain ECM are proteoglycans, hyaluronan, link-proteins like TN-C, and others (Figure 1) (35). Another ECM type in the brain is the basement membrane that covers blood vessels and is part of the perivascular space. Deregulated ECM dynamics is a hallmark of cancer. The ECM of glioma differs from that of the healthy brain. Whereas universal ECM components are expressed uniformly in healthy brains (36), in high-grade glioma fibrous proteins and laminin are upregulated (15, 37). Besides, the interaction of the ECM component hyaluronan with its receptor CD44, both being overexpressed in glioma cells, is a major requirement for glioma invasion (3840). For glioma cells to invade the healthy brain tissue, the intact ECM has to be destroyed and remodeled. ECM degrading and remodeling enzymes include several MMPs, A Disintegrin and Metalloproteinase with Thrombospondin Motifs (ADAMTS), the serine protease plasmin, 6-O-sulfatases, heparanases, cathepsins, and urokinase (uPa). These enzymes are not only regulated at the transcriptional and translational levels but also post-translationally by their functionally inhibitory pro-domains or by selective natural proteinase inhibitors (41).

Fig 1

Figure 1 Mechanisms involved in the migration and invasion of glioblastoma (GBM). The migrational phenotype of GBM cells is regulated by a complex interplay of different factors, signaling cascades, as well as cellular and environmental features.

MATRIX-METALLOPROTEINASES

MMPs are a family of secreted or membrane-anchored endoproteinases (42). Their main function is the degradation and remodeling of the ECM. MMP expression in the normal brain is low. In glioma, MMPs are overexpressed or activated. MMP-2 and MMP-9 are of interest for invasive processes in gliomas as their expression correlates with tumor grade and progression (43, 44). MMP-2 and MMP-9 convert latent pro-migratory transforming growth factor (TGF)-β into its active form, which in turn induces MMP-2 in a feedback loop (see the section “The Role of TGF-β in Glioma Cell Motility” (4547)). MMP-9 expression or activity can be regulated by: activation of signal transducer and activator of transcription (STAT)3; epidermal growth factor (EGF); FN; vitronectin (VN); interleukin (IL)-1β; tumor necrosis factor (TNF)-α; and TGF-β (4752). Furthermore, glioma cells exploit MMP-14 that is expressed by surrounding microglia cells (53). MMP-14 activates MMP-2 by cleaving its pro-peptide (54, 55). Furthermore, MMP-3, -7, -12, -13, -16, -19, and -26 are also highly expressed and mostly associated with enhanced glioma invasion (5663). MMPs are inhibited by the four tissue inhibitors of metalloproteinases (TIMP), TIMP-1–4. They inhibit all MMPs but also have other functions including MMP activation. TIMP-2 can form a ternary complex with pro-MMP-2 and MMP-14 that is necessary for efficient MMP-2 activation (55, 64). High TIMP-1 levels and TIMP-3 silencing are associated with a poor prognosis for glioma patients (6568). Due to these paradoxical effects, the important role of TIMPs in glioma invasion remains elusive.

INTEGRINS—THE LINK BETWEEN THE ECM AND CELLS

Integrins are catalytic inactive heterodimeric transmembrane glycoproteins responsible for cell–ECM interactions. They are the link between the ECM and the cytoskeleton and important for signal transduction. To date, 24 integrins composed of different combinations of 18 α- and 8 β-subunits have been identified (69). The α/β combination determines ligand specificity. Typical ECM ligands for integrins are laminin, collagen, and FN, which are part of the basement membrane in the brain and are expressed by high-grade gliomas (70). Other integrin ligands are thrombospondin (TSP), osteopontin (OPN), VN, and TN-C, all being overexpressed in gliomas. Upon ligand binding, integrins form clusters, leading to activation of the focal adhesion kinase (FAK) and finally to enhanced migration (71). FAK is active and overexpressed in gliomas, and its expression correlates with the tumor grade (7274). Upon integrin clustering, the cytoplasmic domain attaches to cytoskeletal components to form focal adhesion points at the leading edge of migrating cells (75). This adhesion points give cells a polarity which enable them to move forward. In GBM, integrin β1 is overexpressed and is associated with migration (76, 77). Integrin α9β1 expression correlates with glioma grade and influences MMP-9 expression (78, 79). Furthermore, integrin α5β1 can stimulate MMP-2 expression upon interaction with angiopoietin (80). In addition, integrin αvβ3 and αvβ5 expression is associated with disease progression. Both can bind to the latency-associated peptide (LAP) of the LAP-TGF-β complex and thereby release active TGF-β (81, 82). In summary, integrins are substantial for glioma cell migration, establishing the link between the brain ECM and the tumor cells (Figure 1).

CHONDROITIN SULFATE PROTEOGLYCANS, GLYCOPROTEINS, AND GALECTINS

One important class of proteoglycans are chondroitin sulfate proteoglycans (CSPG), which are overexpressed in glioma and associated with increased glioma invasion (83). A subgroup of CSPG, the lecticans, forms tertiary complexes with hyaluronan and TN-R. Three of them, versican, BEHAB/brevican, and neurocan, are overexpressed in glioma and enhance glioma motility (8486).

Invasion-promoting ECM glycoproteins secreted in glioma are: Secreted Protein Acidic and Rich in Cysteine (SPARC); TN-C supporting cell adhesion through integrin binding; OPN and VN (8790). In addition, TSP-1, a multifunctional matrix glycoprotein, is implicated in cell adhesion, migration, invasion, and activation of TGF-β (91; see the section “The Role of TGF-β in Glioma Cell Motility”). Galectins are soluble lectins with specificity for β-galactoside which allow them to bind to proteoglycans and glycoproteins in the brain ECM (92). In malignant gliomas, galectin-1, -3, and -8 are overexpressed and promote glioma cell migration and invasion by modulating the actin cytoskeleton (9396).

Migration-Associated Changes of the Cytoskeleton

Cell migration is a multistep process initiated by binding of chemoattractants or pro-migratory factors to cell surface receptors, followed by the activation or inactivation of diverse small GTPases and cytoskeleton reorganization (97). The resulting structures are called filopodia, lamellipodia, and podosomes. Turnover of adhesion site formation at the cell front and disruption at the rear is essential for cell movement (98).

SMALL GTPASES

The most important and well-characterized small GTPases associated with cytoskeletal remodeling are: RhoA, which is responsible for coordination of contractility at the cell body and cell rear; RAC-1 that regulates protrusion formation at the leading edge; and CDC42 that modulates cell polarity (99). RAC-1 protein levels correlate with tumor grade in astrocytomas. In addition, RAC-1 is hyperactivated in GBM (100). Enhanced activity of CDC42 and RAC-1 has been reported in infiltrating glioma cells (101). Migration-associated small GTPase activity is regulated by a variety of factors and signals. Rho GTPase activity is mediated by several receptors and effectors. In GBM, two members of the TNF receptor superfamily act through RAC-1: TNF-like weak inducer of apoptosis (TWEAK) and TNF receptor superfamily member 19 (TROY) (99, 102). EGFRvIII, a truncated and constitutively active EGF receptor, and Platelet Growth Factor Receptor alpha (PDGFRα) activate RAC-1-mediated migration through tyrosine protein kinase SRC-dependent DOCK180 phosphorylation (103, 104). RAC-1 is also activated by the IQ-domain GTPase-Activating Protein (IQGAP)-1/ADP-Ribosylation Factor 6 (ARF6), neurotensin, and ephrinB3 signaling (105107). RAC-1 activity is further modulated by CDC42 (104, 108) as well as by axonal guidance molecules (see the section “Axonal Guidance Molecules”). RhoA activity correlates with increased glioma cell migration. Functional evidence for the role of RhoA has been demonstrated via inhibition of the RhoA effector ROCK, which leads to enhanced invasion due to the fact that ROCK, together with mDia, coordinates stress fiber formation and focal adhesion, thereby exacerbating migration. The activity of Rho and RAC GTPases is tightly regulated by three main proteins: guanine nucleotide exchange factors (GEFs), GTPase-activating proteins (GAPs), and guanine nucleotide dissociation inhibitors. Many GEFs (e.g., Ect2, ARHGEF7 [βPIX], SWAP, SGEF, Vav3, Trio, Dock180, and Dock9) have been correlated with glioma pathology, higher tumor grade, and glioma invasion, in particular when co-localized with small GTPases (99).

ACTIN REARRANGEMENT, ADHESION COMPLEXES, AND CELLULAR PROTRUSIONS

Nonmotile cells show nonpolarized cell morphology. In these cells, the machinery for actin filament and protrusions formation is inactive. Protrusion formation and actin polymerization requires, besides actin, at least six other proteins: the Arp2/3 complex; an Arp2/3 complex-activating nucleation promoting factor (NPF); a barbed-end capping protein; cofilin and profilin, the latter binding both ADP-bound and ATP-bound actin monomers (109). Lamellipodia are flat, branched, sheet-like actin membrane protrusions that drive cell migration by attaching to the substrate and generating force at the leading edge. Filopodia are thin, finger-like projections beyond the lamellipodial edge, composed of long, bundled, and unbranched actin filaments. No Arp2/3 complex or cofilin are present in filopodia. Invadopodia/podosomes are ventral membrane protrusions responsible for ECM degradation with a not yet well-characterized actin organization (98).

The Wiskott–Aldrich Syndrome (WASP) family consists of two principal classes of proteins: WASPs and SCAR/WAVEs. WASP/N-WASP induces invadopodia and podosome formation, while WAVEs are key regulators of lamellipodia. Cofilin, involved in de-polymerization and polymerization of actin filaments, is highly expressed in migrating GBM cells. It is phosphorylated and inactivated by LIM1/2 kinase. For proper migration and protrusion formation, cofilin and LIM kinase activity must be perfectly balanced. Invadopodia formation is dependent on the activity of cortactin, an actin-binding protein (98).

During cell movement, focal adhesion complexes (FACs) are formed to connect the rearranged actin cytoskeleton to the ECM. While integrin clustering is the first step for FAC formation, microtubule extension promotes FAC disruption. Several studies reported a transport of integrins from the rear to the front of the cell during migration, maintaining the focal adhesion turnover. The presence of large focal adhesions creates more links to actin stress fibers and makes cell movement more difficult (110). The molecular structure of FAC includes integrins, intracellularly bound to paxillin and talin, which subsequently recruit FAK and vinculin. FAK then phosphorylates alpha-actinin, leading to cross-links with actin filaments. The resulting structures lead to alterations of the cell morphology and the generation of traction force necessary to move the cell body. Recent reports indicate that focal adhesion protein expression, like talin and alpha-actinin, is related to the invasiveness of glioma cells (111).

Ion Channels and Their Contribution to Glioma Cell Migration

AUTOCRINE GLUTAMATE SIGNALING

Gliomas express glutamate receptors (GluRs) like α-amino-3-hydroxy-5-methylisoxazole-4-propionic acid receptors (AMPAR), N-methyl-D-aspartate (NMDA) receptors, and metabotropic mGluRs. AMPARs are composed of four types of subunits: GluR1–4. Through autocrine glutamate signaling, they contribute to enhanced glioma cell invasion (112, 113). The subunits, especially GluR2, influence the cation permeability of AMPAR. In the presence of GluR2, the channel is Ca2+ impermeable, the situation in the mature and healthy brain (114, 115). In glioma, GluR2 is not expressed, leading to high Ca2+ permeability (116, 117). Artificial GluR2 overexpression in glioma cells inhibits migration (117, 118). Overexpression of GluR1 positively correlates with glioma cell adhesion to collagen, whereas stimulation of AMPAR leads to detachment from the ECM. In a mouse glioma model, overexpression of GluR1 results in enhanced invasion of glioma cells into the perivascular space similar to patterns described in human GBM.

HYDRODYNAMIC MODEL OF GLIOMA CELL MIGRATION

Glioma cells migrate through the extracellular space in the brain. To aid such migration, they reduce their volume by more than 30% by releasing cytoplasmic water (119). For this purpose, glioma cells exploit ion channels which normally function as membrane potential regulators (Figure 1). Unlike adult neurons, glioma cells have high intracellular Cl levels (120). This is due to the constitutive expression and prolonged activity of the Na+/K+/Cl cotransporter 1 (NKCC1) that correlates with glioma grade and invasiveness (121). Upon opening of Cl channels, the outflow of Cl is accompanied by the efflux of water through aquaporins due to osmotic forces, leading to volume shrinkage. In glioma, the chloride channels ClC-2 and ClC-3 are functionally expressed, and blocking them reduces glioma migration (122124).

The K+ gradient, regulated by Na+/K+-ATPase, is essential for invasion (125). The KCa family of Ca2+-activated K+ channels, especially KCa3.1, is overexpressed in 32% of the glioma patients, and its expression correlates with patient survival (126). KCa3.1 is localized at the leading edge of migrating cells, and its inhibition results in reduced migration (127, 128). The bradykinin receptor B2 (B2R) is also expressed at the leading edge of migrating glioma cells. It is a critical attractor of glioma cells toward the vasculature, and an activator of ion channels (127, 129). Binding of bradykinin to B2R leads to increases in intracellular Ca2+ which induces the opening of the KCa3.1 and ClC-3 channels, resulting in the efflux of Cl, K+, and water (16, 127, 130). As a result, the glioma cells shrink which enable them to migrate through the narrow space of the brain.

Axonal Guidance Molecules

Glioma cell movement can also occur along myelinated neuronal axons of white matter tracts. A multitude of proteins act as axonal guidance molecules by either attracting or repelling axonal growth cones and modulating neural cell motility during development (Figure 1). The most prominent axonal guidance molecules are: ephrins (Eph); netrins; Slits and their roundabout (Robo) receptors; semaphorins (Sema) and their receptors plexin and neuropilin (NRP) (131).

EPHRINS

Ephrins serve as ligands of ephrin receptors (EphRs), a family of proteins containing nine EphR class A and five EphR class B members. Interaction of Eph and EphR regulates cell–cell interaction by forward (Eph to EphR) or reverse (EphR to Eph) signaling. Eph regulates cell migration, adhesion, morphology, differentiation, proliferation, and survival through Jun-N-terminal kinase (JNK), STAT3, PKB/AKT, Rho GTPase, and paxillin pathways. Recent studies have detected an abnormal expression of EphB1 receptors in brain tumors (132). Eph proteins have a dual role in glioma cell migration: negative regulation that inhibits migration and positive regulation that promotes migration (133, 134). Therefore, it could be postulated that these proteins might serve as regulators of the “Go or Grow” behavior of GBM.

NETRINS AND SLIT/ROBO

Netrins are a family of laminin-related proteins. Netrin-1, the most prominent representative of the netrin family, is widely expressed in fetal and adult brain tissues. Its expression is associated with progression of various types of human cancers. Netrin-1 binds to UNC5-family dependence receptor (DR) deleted in colorectal cancer (DCC), or other UNC5 molecules. While the absence of netrin-1, DCC/UNC induces apoptosis, the absence of the DRs or enhanced netrin-1 expression is tumorigenic. Netrin expression is associated with poor patient prognosis in lower grade gliomas. In GBM cells, elevated netrin expression activates notch signaling, finally resulting in the gain of stemness and enhancement of invasiveness of these cells (135).

Slit (Slit 1–3) and the Robo receptor family proteins are evolutionarily conserved molecules. During normal development, secreted Slit proteins regulate axon guidance and neuronal precursor cell migration by mediating chemo-repulsive signals on cells expressing Robo. In glioma, Slit2 and Robo1 provide different patterns. By hypermethylation of its promoter, the expression of Slit is low in most gliomas (136), whereas the expression of Robo1 is high. Slit2/Robo1 signaling inhibits glioma cell migration and invasion by inactivation of CDC42 signaling. In vivo, Slit-2 mitigates infiltration of glioma cells into the healthy brain (137), indicating that a chemo-repulsive signal transmitted by the interaction of Slit2/Robo1 participates in glioma cell migration or guidance (138).

SEMAPHORINS AND THEIR RECEPTORS

Semaphorins (Sema), originally identified as guidance molecules that navigate axon growth in the brain, fall into eight subclasses of secreted, membrane-anchored, and transmembrane proteins (139). Class 3 semaphorins (Sema3) transfer their function through a receptor complex consisting of plexins and neuropilin (NRP)-1 and -2 (140, 141). Downstream signaling of Sema involves RhoA, RAC-1, and cofilin, leading to the reorganization of the cytoskeleton (142). In GBM cells, inactivation of RhoA by Sema3F leads to the collapse of the cytoskeleton, whereas inhibition of Sema3F promotes cell motility (143, 144). Similar effects have been observed for Sema3G (145), and higher expression of Sema3G in GBM patients has been associated with a better prognosis (146). While Sema3A, 3B, and 3F show antitumorigenic properties in many cancers, other Sema3 family members are associated with tumor progression. Overexpression of Sema3C promotes cell invasion of prostate cancer cell lines, whereas enhanced expression of Sema3E induces metastasis in lung cancer (147, 148). Regarding this dual function of semaphorins, it should be kept in mind that the signaling complexes of Semas and their Robo receptors as well as the downstream signaling cascades that are modulated by Semas are complex and interconnected, which then might finally determine whether they work in a pro- or anti-migratory fashion.

The Role of TGF-β in Glioma Cell Motility

The TGF-β superfamily of cytokines consists of TGF-β 1–3 which are master regulators of inflammation and cell differentiation. They play a key role in tumor progression and metastasis (149). After binding to the TGF-β receptor (TGFβ-R)-I, TGFβ-RII is phosphorylated. This in turn phosphorylates SMAD2/3, which then combines with SMAD4. This complex translocates to the nucleus and regulates gene expression (150). TGF-β is heavily secreted by glioma cells in vitro and in vivo. TGF-β promotes a mesenchymal phenotype in GBM cells, enhancing invasion and migration in vitro, and in an orthotopic mouse model (151). TGF-β also stimulates the production of reactive oxygen species (ROS), and activates ERK1/2, JNK, and NFκB. NFκB finally upregulates the expression of MMP-9 (152). Other mechanisms of TGF-β influencing the ECM and promoting migration include the upregulation of integrin αvβ3 and the versican isoforms V0/V1 (84, 153). Furthermore, TGF-β suppresses phosphatase and tensin homolog (PTEN) in glioma cells through enhanced miR10a/b expression (154). In patient samples, TGFB1I1 (TGF-β1-induced transcript 1) expression was found to be correlated with tumor grade, and activation of EMT pathways (152). In reaction to radiation treatment, the invasion capability of glioma cells is enhanced and TGF-β is upregulated. This suggests a role for TGF-β in treatment resistance (155).

EMT-Like Processes

EMT is a process by which epithelial cells lose their polarity and cell–cell adhesion, resulting in a mesenchymal phenotype characterized by enhanced motility, chemo-resistance, and stem-like properties. EMT is involved in various biological functions such as wound healing, embryonic development, and fibrosis (156). In epithelial carcinoma, EMT is a well-established driver of invasion and metastasis (157), and even though gliomas are nonepithelial tumors, EMT-like processes have been described (158). Among the signals that have been shown to induce EMT in glioma are TGF-β, EGF, and Hypoxia-Inducible Factor (HIF; Figure 1) (159).

TWIST, SNAIL, SLUG, AND ZEB

TWIST1 and TWIST2 are helix-loop-helix TFs involved in EMT during development and cancer progression (160). In glioma, TWIST was found to be a possible prognostic marker, and its expression correlates with tumor grade (161, 162). TWIST overexpression promotes invasion of glioma cells in vitro and in orthotopic glioma xenotransplants in vivo by inducing the expression of EMT-associated genes like MMP-2 and FN-1. The SNAIL family of transcriptional repressors consisting of SNAIL/SNAI1 and SLUG/SNAI2 is known to drive invasion and metastasis in various carcinomas (163). SNAIL binds to E-box DNA sequences of genes related to an epithelial phenotype through carboxy-terminal zinc-finger domains, thereby suppressing their expression. Knockdown of SNAIL in glioma cells by siRNA diminished glioma migration and invasion (164, 165). In GBM, the Rho family GTPase (RND)-3 has been shown to promote the degradation of SNAIL in vitro and in vivo, while downregulation of RND3 strongly induces SNAIL expression and migration (166). SLUG expression was found to correlate with histologic grade and invasive phenotype in glioma, whereas knockdown of SLUG attenuated invasion and prolonged survival in an intracranial mouse model (167).

The TFs Zinc-finger E-box Binding homeobox proteins (ZEB)-1 and -2 also bind to E-boxes of DNA sequences, thereby repressing cell polarity-associated genes such as E-cadherin/CDH1, cell–adhesion molecules, and stemness-inhibiting miR-200 (168, 169). In GBM patients, ZEB-1 overexpression correlated with poor overall survival. Glioma cells implanted in mice brain were less invasive after knockdown of ZEB-1. ZEB-1 and PDGFRα were found to be co-expressed in tissue samples from GBM patients, while high expression of both ZEB-1 and activated PDGFRα was identified to significantly coincide with poor survival. The same study further established Protein Tyrosine Phosphatase/Nonreceptor type (PTPN)-1 as a regulator of ZEB-1-induced and PDGFR-induced EMT in glioma (170). EMT may also be directly promoted by the microenvironment of GBM. Both the hypoxic marker HIF1α and ZEB-1 were shown to colocalize in hypoxic areas of human GBM. In glioma cells, the suppression of HIF1α negatively affected the level of ZEB-1 (22). ZEB-2 was overexpressed in glioma tissue samples compared to healthy brain tissue, and higher expression of ZEB-2 correlated with glioma pathology grading. Knockdown of ZEB-2 showed an upregulation of E-cadherin, whereas N-cadherin and SNAIL were repressed (171).

CADHERINS

Cadherins are Ca2+-dependent transmembrane molecules with an important role in cell to cell adhesion, recognition, and signaling (172). In epithelial cancers, the loss of E-cadherin and an increased expression of N-cadherin, the so-called “cadherin switch,” is considered to be a hallmark of EMT (173). In tissues of GBM and healthy brain, the expression of E-cadherin is generally only marginal (174, 175). However, in a minor subset of GBM showing epithelial differentiation, high expression of E-cadherin is observed, correlating with poorer clinical outcome compared to GBM with low or no E-cadherin expression. Glioma cells with high E-cadherin expression show greater invasion when orthotopically implanted in mice (176). In contrast to its role in carcinoma, N-cadherin is frequently downregulated in GBM compared to the healthy brain (177, 178). N-cadherin overexpression has been shown to decrease glioma invasion in vitro and in vivo (179). Interestingly, the role of N-cadherin in glioma is postulated not only to be determined by its expression level but also by its distribution in the cell membrane (180). ZEB-1 knockdown in GBM cells showed a loss of invasiveness and concentration of N-cadherin to the juxtaposed membranes between adjacent cells; the axon-guidance molecule Robo-1 mediated by ZEB-1 can reverse this process by severing the anchorage of N-cadherin to the cytoskeleton (181).

Conclusion

Migration and invasion of glioma cells in the brain follow different migratory routes. It is a complex process regulated by the surrounding environmental conditions, and interconnected by diverse signaling cascades. Understanding the process of migration and invasion of glioma cells is of central importance since these characteristics make GBM aggressive and complete resection impossible. Identifying the molecular mechanisms that govern the motility of GBM cells will help develop new therapeutic strategies to treat this deadly tumor.

Conflict of interest: The authors declare no potential conflicts of interest with respect to research, authorship, and/or publication of this manuscript.

Copyright and permission statement: To the best of our knowledge, the materials included in this chapter do not violate copyright laws. All original sources have been appropriately acknowledged and/or referenced. Where relevant, appropriate permissions have been obtained from the original copyright holder(s).

REFERENCES

  1. Johansen MD, Rochat P, Law I, Scheie D, Poulsen HS, Muhic A. Presentation of two cases with early extracranial metastases from glioblastoma and review of the literature. Case Rep Oncol Med. 2016;2016:8190950. http://dx.doi.org/10.1155/2016/8190950
  2. Demuth T, Berens ME. Molecular mechanisms of glioma cell migration and invasion. J Neuro-Oncol. 2004;70(2):217–28. http://dx.doi.org/10.1007/s11060-004-2751-6
  3. Altieri R, Zenga F, Fontanella MM, Cofano F, Agnoletti A, Spena G, et al. Glioma surgery: Technological advances to achieve a maximal safe resection. Surg Technol Int. 2015;27:297–302.
  4. Cuddapah VA, Robel S, Watkins S, Sontheimer H. A neurocentric perspective on glioma invasion. Nat Rev Neurosci. 2014;15(7):455–65. http://dx.doi.org/10.1038/nrn3765
  5. Kathagen-Buhmann A, Schulte A, Weller J, Holz M, Herold-Mende C, Glass R, et al. Glycolysis and the pentose phosphate pathway are differentially associated with the dichotomous regulation of glioblastoma cell migration versus proliferation. Neuro Oncol. 2016;18(9):1219–29. http://dx.doi.org/10.1093/neuonc/now024
  6. Höring E, Harter PN, Seznec J, Schittenhelm J, Bühring HJ, Bhattacharyya S, et al. The “go or grow” potential of gliomas is linked to the neuropeptide processing enzyme carboxypeptidase E and mediated by metabolic stress. Acta Neuropathol. 2012;124(1):83–97. http://dx.doi.org/10.1007/s00401-011-0940-x
  7. Lu DY, Leung YM, Cheung CW, Chen YR, Wong KL. Glial cell line-derived neurotrophic factor induces cell migration and matrix metalloproteinase-13 expression in glioma cells. Biochem Pharmacol. 2010;80(8):1201–9. http://dx.doi.org/10.1016/j.bcp.2010.06.046
  8. Gladson CL. The extracellular matrix of gliomas: Modulation of cell function. J Neuropathol Exp Neurol. 1999;58(10):1029–40. http://dx.doi.org/10.1097/00005072-199910000-00001
  9. Bougnaud S, Golebiewska A, Oudin A, Keunen O, Harter PN, Mader L, et al. Molecular crosstalk between tumour and brain parenchyma instructs histopathological features in glioblastoma. Oncotarget. 2016;7(22):31955–71. http://dx.doi.org/10.18632/oncotarget.7454
  10. da Fonseca AC, Badie B. Microglia and macrophages in malignant gliomas: Recent discoveries and implications for promising therapies. Clin Dev Immunol. 2013;2013:264124.
  11. Rao JS. Molecular mechanisms of glioma invasiveness: The role of proteases. Nat Rev Cancer. 2003;3(7):489–501. http://dx.doi.org/10.1038/nrc1121
  12. Lefranc F, Brotchi J, Kiss R. Possible future issues in the treatment of glioblastomas: Special emphasis on cell migration and the resistance of migrating glioblastoma cells to apoptosis. J Clin Oncol. 2005;23(10):2411–22. http://dx.doi.org/10.1200/JCO.2005.03.089
  13. Lo CM, Wang HB, Dembo M, Wang YL. Cell movement is guided by the rigidity of the substrate. Biophys J. 2000;79(1):144–52. http://dx.doi.org/10.1016/S0006-3495(00)76279-5
  14. Ulrich TA, de Juan Pardo EM, Kumar S. The mechanical rigidity of the extracellular matrix regulates the structure, motility, and proliferation of glioma cells. Cancer Res. 2009;69(10):4167–74. http://dx.doi.org/10.1158/0008-5472.CAN-08-4859
  15. Mahesparan R, Read TA, Lund-Johansen M, Skaftnesmo KO, Bjerkvig R, Engebraaten O. Expression of extracellular matrix components in a highly infiltrative in vivo glioma model. Acta Neuropathol. 2003;105(1):49–57.
  16. Montana V, Sontheimer H. Bradykinin promotes the chemotactic invasion of primary brain tumors. J Neurosci. 2011;31(13):4858–67. http://dx.doi.org/10.1523/JNEUROSCI.3825-10.2011
  17. Yadav VN, Zamler D, Baker GJ, Kadiyala P, Erdreich-Epstein A, DeCarvalho AC, et al. CXCR4 increases in-vivo glioma perivascular invasion, and reduces radiation induced apoptosis: A genetic knockdown study. Oncotarget. 2016;7(50):83701–19.
  18. Brat DJ, Van Meir EG. Vaso-occlusive and prothrombotic mechanisms associated with tumor hypoxia, necrosis, and accelerated growth in glioblastoma. Lab Invest. 2004;84(4):397–405. http://dx.doi.org/10.1038/labinvest.3700070
  19. Szalad A, Katakowski M, Zheng X, Jiang F, Chopp M. Transcription factor Sp1 induces ADAM17 and contributes to tumor cell invasiveness under hypoxia. J Exp Clin Cancer Res. 2009;28:129. http://dx.doi.org/10.1186/1756-9966-28-129
  20. Wang Y, Liu T, Yang N, Xu S, Li X, Wang D. Hypoxia and macrophages promote glioblastoma invasion by the CCL4-CCR5 axis. Oncol Rep. 2016;36(6):3522–8. http://dx.doi.org/10.3892/or.2016.5171
  21. Le Mercier M, Fortin S, Mathieu V, Kiss R, Lefranc F. Galectins and gliomas. Brain Pathol. 2010;20(1):17–27. http://dx.doi.org/10.1111/j.1750-3639.2009.00270.x
  22. Joseph JV, Conroy S, Pavlov K, Sontakke P, Tomar T, Eggens-Meijer E, et al. Hypoxia enhances migration and invasion in glioblastoma by promoting a mesenchymal shift mediated by the HIF1alpha-ZEB1 axis. Cancer Lett. 2015;359(1):107–16. http://dx.doi.org/10.1016/j.canlet.2015.01.010
  23. Kahlert UD, Suwala AK, Raabe EH, Siebzehnrubl FA, Suarez MJ, Orr BA, et al. ZEB1 promotes invasion in human fetal neural stem cells and hypoxic glioma neurospheres. Brain Pathol. 2015;25(6):724–32. http://dx.doi.org/10.1111/bpa.12240
  24. Esencay M, Sarfraz Y, Zagzag D. CXCR7 is induced by hypoxia and mediates glioma cell migration towards SDF-1alpha. BMC Cancer. 2013;13:347. http://dx.doi.org/10.1186/1471-2407-13-347
  25. Zagzag D, Lukyanov Y, Lan L, Ali MA, Esencay M, Mendez O, et al. Hypoxia-inducible factor 1 and VEGF upregulate CXCR4 in glioblastoma: Implications for angiogenesis and glioma cell invasion. Lab Invest. 2006;86(12):1221–32. http://dx.doi.org/10.1038/labinvest.3700482
  26. Giese A, Loo MA, Tran N, Haskett D, Coons SW, Berens ME. Dichotomy of astrocytoma migration and proliferation. Int J Cancer. 1996;67(2):275–82. http://dx.doi.org/10.1002/(SICI)1097-0215(19960717)67:2%3C275::AID-IJC20%3E3.0.CO;2-9
  27. Hatzikirou H, Basanta D, Simon M, Schaller K, Deutsch A. “Go or grow”: The key to the emergence of invasion in tumour progression? Math Med Biol. 2012;29(1):49–65. http://dx.doi.org/10.1093/imammb/dqq011
  28. Dhruv HD, McDonough Winslow WS, Armstrong B, Tuncali S, Eschbacher J, et al. Reciprocal activation of transcription factors underlies the dichotomy between proliferation and invasion of glioma cells. PLoS One. 2013;8(8):e72134. http://dx.doi.org/10.1371/journal.pone.0072134
  29. Godlewski J, Nowicki MO, Bronisz A, Nuovo G, Palatini J, De Lay M, et al. MicroRNA-451 regulates LKB1/AMPK signaling and allows adaptation to metabolic stress in glioma cells. Mol Cell. 2010;37(5):620–32. http://dx.doi.org/10.1016/j.molcel.2010.02.018
  30. Tan X, Wang S, Yang B, Zhu L, Yin B, Chao T, et al. The CREB-miR-9 negative feedback minicircuitry coordinates the migration and proliferation of glioma cells. PLoS One. 2012;7(11):e49570. http://dx.doi.org/10.1371/journal.pone.0049570
  31. Wild-Bode C, Weller M, Rimner A, Dichgans J, Wick W. Sublethal irradiation promotes migration and invasiveness of glioma cells: Implications for radiotherapy of human glioblastoma. Cancer Res. 2001;61(6):2744–50.
  32. Huber SM, Butz L, Stegen B, Klumpp D, Braun N, Ruth P, et al. Ionizing radiation, ion transports, and radioresistance of cancer cells. Front Physiol. 2013;4:212. http://dx.doi.org/10.3389/fphys.2013.00212
  33. Novak U, Kaye AH. Extracellular matrix and the braIn: Components and function. J Clin Neurosci. 2000;7(4):280–90. http://dx.doi.org/10.1054/jocn.1999.0212
  34. Cragg B. Brain extracellular space fixed for electron microscopy. Neurosci Lett. 1979;15(2–3):301–6. http://dx.doi.org/10.1016/0304-3940(79)96130-5
  35. Dityatev A, Seidenbecher CI, Schachner M. Compartmentalization from the outside: The extracellular matrix and functional microdomains in the brain. Trends Neurosci. 2010;33(11):503–12. http://dx.doi.org/10.1016/j.tins.2010.08.003
  36. Lau LW, Cua R, Keough MB, Haylock-Jacobs S, Yong VW. Pathophysiology of the brain extracellular matrix: A new target for remyelination. Nat Rev Neurosci. 2013;14(10):722–9. http://dx.doi.org/10.1038/nrn3550
  37. Serres E, Debarbieux F, Stanchi F, Maggiorella L, Grall D, Turchi L, et al. Fibronectin expression in glioblastomas promotes cell cohesion, collective invasion of basement membrane in vitro and orthotopic tumor growth in mice. Oncogene. 2014;33(26):3451–62. http://dx.doi.org/10.1038/onc.2013.305
  38. Wiranowska M, Ladd S, Moscinski LC, Hill B, Haller E, Mikecz K, et al. Modulation of hyaluronan production by CD44 positive glioma cells. Int J Cancer. 2010;127(3):532–42. http://dx.doi.org/10.1002/ijc.25085
  39. Park JB, Kwak HJ, Lee SH. Role of hyaluronan in glioma invasion. Cell Adh Migr. 2008;2(3):202–7. http://dx.doi.org/10.4161/cam.2.3.6320
  40. Bouterfa H, Janka M, Meese E, Kerkau S, Roosen K, Tonn JC. Effect of changes in the CD44 gene on tumour cell invasion in gliomas. Neuropathol Appl Neurobiol. 1997;23(5):373–9. http://dx.doi.org/10.1111/j.1365-2990.1997.tb01311.x
  41. Lu P, Weaver VM, Werb Z. The extracellular matrix: A dynamic niche in cancer progression. J Cell Biol. 2012;196(4):395–406. http://dx.doi.org/10.1083/jcb.201102147
  42. Konnecke H, Bechmann I. The role of microglia and matrix metalloproteinases involvement in neuroinflammation and gliomas. Clin Dev Immunol. 2013;2013:914104. http://dx.doi.org/10.1155/2013/914104
  43. Forsyth PA, Wong H, Laing TD, Rewcastle NB, Morris DG, Muzik H, et al. Gelatinase-A (MMP-2), gelatinase-B (MMP-9) and membrane type matrix metalloproteinase-1 (MT1-MMP) are involved in different aspects of the pathophysiology of malignant gliomas. Br J Cancer. 1999;79(11–12):1828–35. http://dx.doi.org/10.1038/sj.bjc.6990291
  44. Wang M, Wang T, Liu S, Yoshida D, Teramoto A. The expression of matrix metalloproteinase-2 and -9 in human gliomas of different pathological grades. Brain Tumor Pathol. 2003;20(2):65–72. http://dx.doi.org/10.1007/BF02483449
  45. Wick W, Platten M, Weller M. Glioma cell invasion: Regulation of metalloproteinase activity by TGF-beta. J Neurooncol. 2001;53(2):177–85.
  46. Imai K, Hiramatsu A, Fukushima D, Pierschbacher MD, Okada Y. Degradation of decorin by matrix metalloproteinases: Identification of the cleavage sites, kinetic analyses and transforming growth factor-beta1 release. Biochem J. 1997;322(Pt 3):809–14. http://dx.doi.org/10.1042/bj3220809
  47. Yu Q, Stamenkovic I. Cell surface-localized matrix metalloproteinase-9 proteolytically activates TGF-beta and promotes tumor invasion and angiogenesis. Genes Dev. 2000;14(2):163–76.
  48. Natesh K, Bhosale D, Desai A, Chandrika G, Pujari R, Jagtap J, et al. Oncostatin-M differentially regulates mesenchymal and proneural signature genes in gliomas via STAT3 signaling. Neoplasia. 2015;17(2):225–37. http://dx.doi.org/10.1016/j.neo.2015.01.001
  49. Beliveau A, Mott JD, Lo A, Chen EI, Koller AA, Yaswen P, et al. Raf-induced MMP9 disrupts tissue architecture of human breast cells in three-dimensional culture and is necessary for tumor growth in vivo. Genes Dev. 2010;24(24):2800–11. http://dx.doi.org/10.1101/gad.1990410
  50. Milner R, Crocker SJ, Hung S, Wang X, Frausto RF, del Zoppo GJ. Fibronectin- and vitronectininduced microglial activation and matrix metalloproteinase-9 expression is mediated by integrins alpha5beta1 and alphavbeta5. J Immunol. 2007;178(12):8158–67. http://dx.doi.org/10.4049/jimmunol.178.12.8158
  51. Lin CC, Kuo CT, Cheng CY, Wu CY, Lee CW, Hsieh HL, et al. IL-1 beta promotes A549 cell migration via MAPKs/AP-1- and NF-kappaB-dependent matrix metalloproteinase-9 expression. Cell Signal. 2009;21(11):1652–62. http://dx.doi.org/10.1016/j.cellsig.2009.07.002
  52. Esteve PO, Chicoine E, Robledo O, Aoudjit F, Descoteaux A, Potworowski EF, et al. Protein kinase C-zeta regulates transcription of the matrix metalloproteinase-9 gene induced by IL-1 and TNF-alpha in glioma cells via NF-kappa B. J Biol Chem. 2002;277(38):35150–5. http://dx.doi.org/10.1074/jbc.M108600200
  53. Markovic DS, Vinnakota K, Chirasani S, Synowitz M, Raguet H, Stock K, et al. Gliomas induce and exploit microglial MT1-MMP expression for tumor expansion. Proc Natl Acad Sci USA. 2009; 106(30):12530–5. http://dx.doi.org/10.1073/pnas.0804273106
  54. Sato H, Takino T, Okada Y, Cao J, Shinagawa A, Yamamoto E, et al. A matrix metalloproteinase expressed on the surface of invasive tumour cells. Nature. 199;370(6484):61–5.
  55. Murphy G, Stanton H, Cowell S, Butler G, Knauper V, Atkinson S, et al. Mechanisms for pro matrix metalloproteinase activation. APMIS. 1999;107(1):38–44. http://dx.doi.org/10.1111/j.1699-0463.1999.tb01524.x
  56. Sarkar S, Nuttall RK, Liu S, Edwards DR, Yong VW. Tenascin-C stimulates glioma cell invasion through matrix metalloproteinase-12. Cancer Res. 2006;66(24):11771–80. http://dx.doi.org/10.1158/00085472.CAN-05-0470
  57. Wang J, Li Y, Wang J, Li C, Yu K, Wang Q. Increased expression of matrix metalloproteinase-13 in glioma is associated with poor overall survival of patients. Med Oncol. 2012;29(4):2432–7. http://dx.doi.org/10.1007/s12032-012-0181-4
  58. Yeh WL, Lu DY, Lee MJ, Fu WM. Leptin induces migration and invasion of glioma cells through MMP-13 production. Glia. 2009;57(4):454–64. http://dx.doi.org/10.1002/glia.20773
  59. Wang H, Li XT, Wu C, Wu ZW, Li YY, Yang TQ, et al. miR-132 can inhibit glioma cells invasion and migration by target MMP16 in vitro. Onco Targets Ther. 2015;8:3211–18.
  60. Laurent M, Martinerie C, Thibout H, Hoffman MP, Verrecchia F, Le Bouc Y, et al. NOVH increases MMP3 expression and cell migration in glioblastoma cells via a PDGFR-alpha-dependent mechanism. FASEB J. 2003;17(13):1919–21.
  61. Deng Y, Li W, Li Y, Yang H, Xu H, Liang S, et al. Expression of Matrix Metalloproteinase-26 promotes human glioma U251 cell invasion in vitro and in vivo. Oncol Rep. 2010;23(1):69–78.
  62. Rome C, Arsaut J, Taris C, Couillaud F, Loiseau H. MMP-7 (matrilysin) expression in human brain tumors. Mol Carcinog. 2007;46(6):446–52. http://dx.doi.org/10.1002/mc.20293
  63. Lettau I, Hattermann K, Held-Feindt J, Brauer R, Sedlacek R, Mentlein R. Matrix metalloproteinase-19 is highly expressed in astroglial tumors and promotes invasion of glioma cells. J Neuropathol Exp Neurol. 2010;69(3):215–23. http://dx.doi.org/10.1097/NEN.0b013e3181ce9f67
  64. Jackson HW, Defamie V, Waterhouse P, Khokha R. TIMPs: Versatile extracellular regulators in cancer. Nat Rev Cancer. 2017;17(1):38–53. http://dx.doi.org/10.1038/nrc.2016.115
  65. Crocker M, Ashley S, Giddings I, Petrik V, Hardcastle A, Aherne W, et al. Serum angiogenic profile of patients with glioblastoma identifies distinct tumor subtypes and shows that TIMP-1 is a prognostic factor. Neuro Oncol. 2011;13(1):99–108. http://dx.doi.org/10.1093/neuonc/noq170
  66. Aaberg-Jessen C, Christensen K, Offenberg H, Bartels A, Dreehsen T, Hansen S, et al. Low expression of tissue inhibitor of metalloproteinases-1 (TIMP-1) in glioblastoma predicts longer patient survival. J Neurooncol. 2009;95(1):117–28. http://dx.doi.org/10.1007/s11060-009-9910-8
  67. Nakamura M, Ishida E, Shimada K, Kishi M, Nakase H, Sakaki T, et al. Frequent LOH on 22q12.3 and TIMP-3 inactivation occur in the progression to secondary glioblastomas. Lab Invest. 2005;85(2):165–75. http://dx.doi.org/10.1038/labinvest.3700223
  68. Zhang L, Wang M, Wang W, Mo J. Incidence and prognostic value of multiple gene promoter methylations in gliomas. J Neurooncol. 2014;116(2):349–56. http://dx.doi.org/10.1007/s11060-013-1301-5
  69. Desgrosellier JS, Cheresh DA. Integrins in cancer: Biological implications and therapeutic opportunities. Nat Rev Cancer. 2010;10(1):9–22. http://dx.doi.org/10.1038/nrc2748
  70. D’Abaco GM, Kaye AH. Integrins: Molecular determinants of glioma invasion. J Clin Neurosci. 2007;14(11):1041–8. http://dx.doi.org/10.1016/j.jocn.2007.06.019
  71. Mitra SK, Schlaepfer DD. Integrin-regulated FAK-Src signaling in normal and cancer cells. Curr Opin Cell Biol. 2006;18(5):516–23. http://dx.doi.org/10.1016/j.ceb.2006.08.011
  72. Wang D, Grammer JR, Cobbs CS, Stewart JE, Jr., Liu Z, Rhoden R, et al. p125 focal adhesion kinase promotes malignant astrocytoma cell proliferation in vivo. J Cell Sci. 2000;113(Pt 23):4221–30.
  73. Hecker TP, Grammer JR, Gillespie GY, Stewart J, Jr., Gladson CL. Focal adhesion kinase enhances signaling through the Shc/extracellular signal-regulated kinase pathway in anaplastic astrocytoma tumor biopsy samples. Cancer Res. 2002;62(9):2699–707.
  74. Gutenberg A, Bruck W, Buchfelder M, Ludwig HC. Expression of tyrosine kinases FAK and Pyk2 in 331 human astrocytomas. Acta Neuropathol. 2004;108(3):224–30. http://dx.doi.org/10.1007/s00401-004-0886-3
  75. Ridley AJ, Schwartz MA, Burridge K, Firtel RA, Ginsberg MH, Borisy G, et al. Cell migration: Integrating signals from front to back. Science. 2003;302(5651):1704–9. http://dx.doi.org/10.1126/science.1092053
  76. Gingras MC, Roussel E, Bruner JM, Branch CD, Moser RP. Comparison of cell adhesion molecule expression between glioblastoma multiforme and autologous normal brain tissue. J Neuroimmunol. 1995;57(1–2):143–53. http://dx.doi.org/10.1016/0165-5728(94)00178-Q
  77. Tysnes BB, Larsen LF, Ness GO, Mahesparan R, Edvardsen K, Garcia-Cabrera I, et al. Stimulation of glioma-cell migration by laminin and inhibition by anti-alpha3 and anti-beta1 integrin antibodies. Int J Cancer. 1996;67(6):777–84. http://dx.doi.org/10.1002/(SICI)1097-0215(19960917)67:6%3C777::AID-IJC5%3E3.0.CO;2-O
  78. Brown MC, Staniszewska I, Lazarovici P, Tuszynski GP, Del Valle L, Marcinkiewicz C. Regulatory effect of nerve growth factor in alpha9beta1 integrin-dependent progression of glioblastoma. Neuro Oncol. 2008;10(6):968–80. http://dx.doi.org/10.1215/15228517-2008-0047
  79. Veeravalli KK, Ponnala S, Chetty C, Tsung AJ, Gujrati M, Rao JS. Integrin alpha9beta1-mediated cell migration in glioblastoma via SSAT and Kir4.2 potassium channel pathway. Cell Signal. 2012;24(1):272–81. http://dx.doi.org/10.1016/j.cellsig.2011.09.011
  80. Hu B, Jarzynka MJ, Guo P, Imanishi Y, Schlaepfer DD, Cheng SY. Angiopoietin 2 induces glioma cell invasion by stimulating matrix metalloprotease 2 expression through the alphavbeta1 integrin and focal adhesion kinase signaling pathway. Cancer Res. 2006;66(2):775–83. http://dx.doi.org/10.1158/0008-5472.CAN-05-1149
  81. Ludbrook SB, Barry ST, Delves CJ, Horgan CM. The integrin alphavbeta3 is a receptor for the latency-associated peptides of transforming growth factors beta1 and beta3. Biochem J. 2003;369 (Pt 2):311–18. http://dx.doi.org/10.1042/bj20020809
  82. Bello L, Francolini M, Marthyn P, Zhang J, Carroll RS, Nikas DC, et al. Alpha(v)beta3 and alpha(v) beta5 integrin expression in glioma periphery. Neurosurgery. 2001;49(2):380–9; discussion 90.
  83. Sim H, Hu B, Viapiano MS. Reduced expression of the hyaluronan and proteoglycan link proteins in malignant gliomas. J Biol Chem. 2009;284(39):26547–56. http://dx.doi.org/10.1074/jbc.M109.013185
  84. Onken J, Moeckel S, Leukel P, Leidgens V, Baumann F, Bogdahn U, et al. Versican isoform V1 regulates proliferation and migration in high-grade gliomas. J Neurooncol. 2014;120(1):73–83. http://dx.doi.org/10.1007/s11060-014-1545-8
  85. Viapiano MS, Hockfield S, Matthews RT. BEHAB/brevican requires ADAMTS-mediated proteolytic cleavage to promote glioma invasion. J Neurooncol. 2008;88(3):261–72. http://dx.doi.org/10.1007/s11060-008-9575-8
  86. Varga I, Hutoczki G, Szemcsak CD, Zahuczky G, Toth J, Adamecz Z, et al. Brevican, neurocan, tenascin-C and versican are mainly responsible for the invasiveness of low-grade astrocytoma. Pathol Oncol Res. 2012;18(2):413–20. http://dx.doi.org/10.1007/s12253-011-9461-0
  87. Rempel SA, Golembieski WA, Fisher JL, Maile M, Nakeff A. SPARC modulates cell growth, attachment and migration of U87 glioma cells on brain extracellular matrix proteins. J Neurooncol. 2001;53(2):149–60. http://dx.doi.org/10.1023/A:1012201300188
  88. Brosicke N, Faissner A. Role of tenascins in the ECM of gliomas. Cell Adh Migr. 2015;9(1–2):131–40. http://dx.doi.org/10.1080/19336918.2014.1000071
  89. Jan HJ, Lee CC, Shih YL, Hueng DY, Ma HI, Lai JH, et al. Osteopontin regulates human glioma cell invasiveness and tumor growth in mice. Neuro Oncol. 2010;12(1):58–70. http://dx.doi.org/10.1093/neuonc/nop013
  90. Fukushima Y, Tamura M, Nakagawa H, Itoh K. Induction of glioma cell migration by vitronectin in human serum and cerebrospinal fluid. J Neurosurg. 2007;107(3):578–85. http://dx.doi.org/10.3171/JNS-07/09/0578
  91. Amagasaki K, Sasaki A, Kato G, Maeda S, Nukui H, Naganuma H. Antisense-mediated reduction in thrombospondin-1 expression reduces cell motility in malignant glioma cells. Int J Cancer. 2001;94(4):508–12. http://dx.doi.org/10.1002/ijc.1497
  92. Strik HM, Kolodziej M, Oertel W, Basecke J. Glycobiology in malignant gliomas: Expression and functions of galectins and possible therapeutic options. Curr Pharm Biotechnol. 2012;13(11):2299–307. http://dx.doi.org/10.2174/138920112802502051
  93. Metz C, Doger R, Riquelme E, Cortes P, Holmes C, Shaughnessy R, et al. Galectin-8 promotes migration and proliferation and prevents apoptosis in U87 glioblastoma cells. Biol Res. 2016;49(1):33. http://dx.doi.org/10.1186/s40659-016-0091-6
  94. Toussaint LG, 3rd, Nilson AE, Goble JM, Ballman KV, James CD, Lefranc F, et al. Galectin-1, a gene preferentially expressed at the tumor margin, promotes glioblastoma cell invasion. Mol Cancer. 2012;11:32. http://dx.doi.org/10.1186/1476-4598-11-32
  95. Camby I, Belot N, Lefranc F, Sadeghi N, de Launoit Y, Kaltner H, et al. Galectin-1 modulates human glioblastoma cell migration into the brain through modifications to the actin cytoskeleton and levels of expression of small GTPases. J Neuropathol Exp Neurol. 2002;61(7):585–96. http://dx.doi.org/10.1093/jnen/61.7.585
  96. Debray C, Vereecken P, Belot N, Teillard P, Brion JP, Pandolfo M, et al. Multifaceted role of galectin-3 on human glioblastoma cell motility. Biochem Biophys Res Commun. 2004;325(4):1393–8. http://dx.doi.org/10.1016/j.bbrc.2004.10.181
  97. Mitra SK, Hanson DA, Schlaepfer DD. Focal adhesion kinase: In command and control of cell motility. Nat Rev Mol Cell Biol. 2005;6(1):56–68. http://dx.doi.org/10.1038/nrm1549
  98. Yamaguchi H, Condeelis J. Regulation of the actin cytoskeleton in cancer cell migration and invasion. Biochim Biophys Acta. 2007;1773(5):642–52. http://dx.doi.org/10.1016/j.bbamcr.2006.07.001
  99. Fortin Ensign SP, Mathews IT, Symons MH, Berens ME, Tran NL. Implications of Rho GTPase signaling in glioma cell invasion and tumor progression. Front Oncol. 2013;3:241. http://dx.doi.org/10.3389/fonc.2013.00241
  100. Salhia B, Tran NL, Chan A, Wolf A, Nakada M, Rutka F, et al. The guanine nucleotide exchange factors trio, Ect2, and Vav3 mediate the invasive behavior of glioblastoma. Am J Pathol. 2008;173(6): 1828–38. http://dx.doi.org/10.2353/ajpath.2008.080043
  101. Hirata E, Yukinaga H, Kamioka Y, Arakawa Y, Miyamoto S, Okada T, et al. In vivo fluorescence resonance energy transfer imaging reveals differential activation of Rho-family GTPases in glioblastoma cell invasion. J Cell Sci. 2012;125(Pt 4):858–68. http://dx.doi.org/10.1242/jcs.089995
  102. Paulino VM, Yang Z, Kloss J, Ennis MJ, Armstrong BA, Loftus JC, et al. TROY (TNFRSF19) is overexpressed in advanced glial tumors and promotes glioblastoma cell invasion via Pyk2-Rac1 signaling. Mol Cancer Res. 2010;8(11):1558–67. http://dx.doi.org/10.1158/1541-7786.MCR-10-0334
  103. Feng H, Hu B, Jarzynka MJ, Li Y, Keezer S, Johns TG, et al. Phosphorylation of dedicator of cytokinesis 1 (Dock180) at tyrosine residue Y722 by Src family kinases mediates EGFRvIII-driven glioblastoma tumorigenesis. Proc Natl Acad Sci U S A. 2012;109(8):3018–23. http://dx.doi.org/10.1073/pnas.1121457109
  104. Feng H, Hu B, Liu KW, Li Y, Lu X, Cheng T, et al. Activation of Rac1 by Src-dependent phosphorylation of Dock180(Y1811) mediates PDGFRalpha-stimulated glioma tumorigenesis in mice and humans. J Clin Invest. 2011;121(12):4670–84. http://dx.doi.org/10.1172/JCI58559
  105. Hu B, Shi B, Jarzynka MJ, Yiin JJ, D’Souza-Schorey C, Cheng SY. ADP-ribosylation factor 6 regulates glioma cell invasion through the IQ-domain GTPase-activating protein 1-Rac1-mediated pathway. Cancer Res. 2009;69(3):794–801. http://dx.doi.org/10.1158/0008-5472.CAN-08-2110
  106. Servotte S, Camby I, Debeir O, Deroanne C, Lambert CA, Lapiere CM, et al. The in vitro influences of neurotensin on the motility characteristics of human U373 glioblastoma cells. Neuropathol Appl Neurobiol. 2006;32(6):575–84. http://dx.doi.org/10.1111/j.1365-2990.2006.00760.x
  107. Nakada M, Drake KL, Nakada S, Niska JA, Berens ME. Ephrin-B3 ligand promotes glioma invasion through activation of Rac1. Cancer Res. 2006;66(17):8492–500. http://dx.doi.org/10.1158/00085472.can-05-4211
  108. Fortin SP, Ennis MJ, Schumacher CA, Zylstra-Diegel CR, Williams BO, Ross JT, et al. Cdc42 and the guanine nucleotide exchange factors Ect2 and trio mediate Fn14-induced migration and invasion of glioblastoma cells. Mol Cancer Res. 2012;10(7):958–68. http://dx.doi.org/10.1158/1541-7786.MCR-11-0616
  109. Nicholson-Dykstra S, Higgs HN, Harris ES. Actin dynamics: Growth from dendritic branches. Curr Biol. 2005;15(9):R346–57. http://dx.doi.org/10.1016/j.cub.2005.04.029
  110. Nagano M, Hoshino D, Koshikawa N, Akizawa T, Seiki M. Turnover of focal adhesions and cancer cell migration. Int J Cell Biol. 2012;2012:310616. http://dx.doi.org/10.1155/2012/310616
  111. Sen S, Ng WP, Kumar S. Contributions of talin-1 to glioma cell–matrix tensional homeostasis. J R Soc Interface. 2012;9(71):1311–17. http://dx.doi.org/10.1098/rsif.2011.0567
  112. Lyons SA, Chung WJ, Weaver AK, Ogunrinu T, Sontheimer H. Autocrine glutamate signaling promotes glioma cell invasion. Cancer Res. 2007;67(19):9463–71. http://dx.doi.org/10.1158/00085472.CAN-07-2034
  113. Rojas A, Dingledine R. Ionotropic glutamate receptors: Regulation by G-protein-coupled receptors. Mol Pharmacol. 2013;83(4):746–52. http://dx.doi.org/10.1124/mol.112.083352
  114. Kim DY, Kim SH, Choi HB, Min C, Gwag BJ. High abundance of GluR1 mRNA and reduced Q/R editing of GluR2 mRNA in individual NADPH-diaphorase neurons. Mol Cell Neurosci. 2001;17(6):1025–33. http://dx.doi.org/10.1006/mcne.2001.0988
  115. Isaac JT, Ashby MC, McBain CJ. The role of the GluR2 subunit in AMPA receptor function and synaptic plasticity. Neuron. 2007;54(6):859–71. http://dx.doi.org/10.1016/j.neuron.2007.06.001
  116. Maas S, Patt S, Schrey M, Rich A. Underediting of glutamate receptor GluR-B mRNA in malignant gliomas. Proc Natl Acad Sci U S A. 2001;98(25):14687–92. http://dx.doi.org/10.1073/pnas.251531398
  117. Ishiuchi S, Tsuzuki K, Yoshida Y, Yamada N, Hagimura N, Okado H, et al. Blockage of Ca(2+)permeable AMPA receptors suppresses migration and induces apoptosis in human glioblastoma cells. Nat Med. 2002;8(9):971–8. http://dx.doi.org/10.1038/nm746
  118. Beretta F, Bassani S, Binda E, Verpelli C, Bello L, Galli R, et al. The GluR2 subunit inhibits proliferation by inactivating Src-MAPK signalling and induces apoptosis by means of caspase 3/6-dependent activation in glioma cells. Eur J Neurosci. 2009;30(1):25–34. http://dx.doi.org/10.1111/j.1460-9568.2009.06804.x
  119. Watkins S, Sontheimer H. Hydrodynamic cellular volume changes enable glioma cell invasion. J Neurosci. 2011;31(47):17250–9. http://dx.doi.org/10.1523/JNEUROSCI.3938-11.2011
  120. Habela CW, Ernest NJ, Swindall AF, Sontheimer H. Chloride accumulation drives volume dynamics underlying cell proliferation and migration. J Neurophysiol. 2009;101(2):750–7. http://dx.doi.org/10.1152/jn.90840.2008
  121. Garzon-Muvdi T, Schiapparelli P, ap Rhys C, Guerrero-Cazares H, Smith C, Kim DH, et al. Regulation of brain tumor dispersal by NKCC1 through a novel role in focal adhesion regulation. PLoS Biol. 2012;10(5):e1001320. http://dx.doi.org/10.1371/journal.pbio.1001320
  122. Olsen ML, Schade S, Lyons SA, Amaral MD, Sontheimer H. Expression of voltage-gated chloride channels in human glioma cells. J Neurosci. 2003;23(13):5572–82.
  123. Soroceanu L, Manning TJ, Jr., Sontheimer H. Modulation of glioma cell migration and invasion using Cl(-) and K(+) ion channel blockers. J Neurosci. 1999;19(14):5942–54.
  124. Lui VC, Lung SS, Pu JK, Hung KN, Leung GK. Invasion of human glioma cells is regulated by multiple chloride channels including ClC-3. Anticancer Res. 2010;30(11):4515–24.
  125. Chen D, Song M, Mohamad O, Yu SP. Inhibition of Na+/K+-ATPase induces hybrid cell death and enhanced sensitivity to chemotherapy in human glioblastoma cells. BMC Cancer. 2014;14:716. http://dx.doi.org/10.1186/1471-2407-14-716
  126. Turner KL, Honasoge A, Robert SM, McFerrin MM, Sontheimer H. A proinvasive role for the Ca(2+)-activated K(+) channel KCa3.1 in malignant glioma. Glia. 2014;62(6):971–81. http://dx.doi.org/10.1002/glia.22655
  127. Cuddapah VA, Turner KL, Seifert S, Sontheimer H. Bradykinin-induced chemotaxis of human gliomas requires the activation of KCa3.1 and ClC-3. J Neurosci. 2013;33(4):1427–40. http://dx.doi.org/10.1523/JNEUROSCI.3980-12.2013
  128. D’Alessandro G, Catalano M, Sciaccaluga M, Chece G, Cipriani R, Rosito M, et al. KCa3.1 channels are involved in the infiltrative behavior of glioblastoma in vivo. Cell Death Dis. 2013;4:e773.
  129. Thompson EG, Sontheimer H. A role for ion channels in perivascular glioma invasion. Eur Biophys J. 2016;45(7):635–48. http://dx.doi.org/10.1007/s00249-016-1154-x
  130. Seifert S, Sontheimer H. Bradykinin enhances invasion of malignant glioma into the brain parenchyma by inducing cells to undergo amoeboid migration. J Physiol. 2014;592(22):5109–27. http://dx.doi.org/10.1113/jphysiol.2014.274498
  131. Chedotal A, Kerjan G, Moreau-Fauvarque C. The brain within the tumor: New roles for axon guidance molecules in cancers. Cell Death Differ. 2005;12(8):1044–56. http://dx.doi.org/10.1038/sj.cdd.4401707
  132. Wei W, Wang H, Ji S. Paradoxes of the EphB1 receptor in malignant brain tumors. Cancer Cell Intl. 2017;17:21. http://dx.doi.org/10.1186/s12935-017-0384-z
  133. Holmberg J, Armulik A, Senti KA, Edoff K, Spalding K, Momma S, et al. Ephrin-A2 reverse signaling negatively regulates neural progenitor proliferation and neurogenesis. Genes Dev. 2005;19(4):462–71. http://dx.doi.org/10.1101/gad.326905
  134. Nakada M, Niska JA, Miyamori H, McDonough WS, Wu J, Sato H, et al. The phosphorylation of EphB2 receptor regulates migration and invasion of human glioma cells. Cancer Res. 2004;64(9):3179–85. http://dx.doi.org/10.1158/0008-5472.can-03-3667
  135. Ylivinkka I, Sihto H, Tynninen O, Hu Y, Laakso A, Kivisaari R, et al. Motility of glioblastoma cells is driven by netrin-1 induced gain of stemness. J Exp Clin Cancer Res. 2017;36(1):9. http://dx.doi.org/10.1186/s13046-016-0482-0
  136. Dickinson RE, Dallol A, Bieche I, Krex D, Morton D, Maher ER, et al. Epigenetic inactivation of SLIT3 and SLIT1 genes in human cancers. Br J Cancer. 2004;91(12):2071–8. http://dx.doi.org/10.1038/sj.bjc.6602222
  137. Yiin JJ, Hu B, Jarzynka MJ, Feng H, Liu KW, Wu JY, et al. Slit2 inhibits glioma cell invasion in the brain by suppression of Cdc42 activity. Neuro Oncol. 2009;11(6):779–89. http://dx.doi.org/10.1215/15228517-2009-017
  138. Mertsch S, Schmitz N, Jeibmann A, Geng JG, Paulus W, Senner V. Slit2 involvement in glioma cell migration is mediated by Robo1 receptor. J Neurooncol. 2008;87(1):1–7. http://dx.doi.org/10.1007/s11060-007-9484-2
  139. Tessier-Lavigne M, Goodman CS. The molecular biology of axon guidance. Science. 1996;274(5290):1123–33. http://dx.doi.org/10.1126/science.274.5290.1123
  140. Kong Y, Janssen BJ, Malinauskas T, Vangoor VR, Coles CH, Kaufmann R, et al. Structural basis for plexin activation and regulation. Neuron. 2016;91(3):548–60. http://dx.doi.org/10.1016/j.neuron.2016.06.018
  141. Janssen BJ, Malinauskas T, Weir GA, Cader MZ, Siebold C, Jones EY. Neuropilins lock secreted semaphorins onto plexins in a ternary signaling complex. Nat Struct Mol Biol. 2012;19(12):1293–9. http://dx.doi.org/10.1038/nsmb.2416
  142. Derijck AA, Van Erp S, Pasterkamp RJ. Semaphorin signaling: Molecular switches at the midline. Trends Cell Biol. 2010;20(9):568–76. http://dx.doi.org/10.1016/j.tcb.2010.06.007
  143. Li X, Law JW, Lee AY. Semaphorin 5A and plexin-B3 regulate human glioma cell motility and morphology through Rac1 and the actin cytoskeleton. Oncogene. 2012;31(5):595–610.
  144. Li X, Lee AY. Semaphorin 5A and plexin-B3 inhibit human glioma cell motility through RhoGDIalphamediated inactivation of Rac1 GTPase. J Biol Chem. 2010;285(42):32436–45. http://dx.doi.org/10.1074/jbc.M110.120451
  145. Zhou X, Ma L, Li J, Gu J, Shi Q, Yu R. Effects of SEMA3G on migration and invasion of glioma cells. Oncology Rep. 2012;28(1):269–75.
  146. Karayan-Tapon L, Wager M, Guilhot J, Levillain P, Marquant C, Clarhaut J, et al. Semaphorin, neuropilin and VEGF expression in glial tumours: SEMA3G, a prognostic marker? Br J Cancer. 2008;99(7):1153–60. http://dx.doi.org/10.1038/sj.bjc.6604641
  147. Herman JG, Meadows GG. Increased class 3 semaphorin expression modulates the invasive and adhesive properties of prostate cancer cells. Int J Oncol. 2007;30(5):1231–8. http://dx.doi.org/10.3892/ijo.30.5.1231
  148. Christensen C, Ambartsumian N, Gilestro G, Thomsen B, Comoglio P, Tamagnone L, et al. Proteolytic processing converts the repelling signal Sema3E into an inducer of invasive growth and lung metastasis. Cancer Res. 2005;65(14):6167–77. http://dx.doi.org/10.1158/0008-5472.CAN-04-4309
  149. Massague J. TGF-beta signaling in development and disease. FEBS Lett. 2012;586(14):1833. http://dx.doi.org/10.1016/j.febslet.2012.05.030
  150. Massaous J, Hata A. TGF-beta signalling through the Smad pathway. Trends Cell Biol. 1997;7(5): 187–92. http://dx.doi.org/10.1016/S0962-8924(97)01036-2
  151. Joseph JV, Conroy S, Tomar T, Eggens-Meijer E, Bhat K, Copray S, et al. TGF-beta is an inducer of ZEB1-dependent mesenchymal transdifferentiation in glioblastoma that is associated with tumor invasion. Cell Death Dis. 2014;5:e1443. http://dx.doi.org/10.1038/cddis.2014.395
  152. Hsieh HL, Wang HH, Wu WB, Chu PJ, Yang CM. Transforming growth factor-beta1 induces matrix metalloproteinase-9 and cell migration in astrocytes: Roles of ROS-dependent ERK- and JNK-NFkappaB pathways. J Neuroinflammation. 2010;7:88. http://dx.doi.org/10.1186/1742-2094-7-88
  153. Platten M, Wick W, Wild-Bode C, Aulwurm S, Dichgans J, Weller M. Transforming growth factors beta(1) (TGF-beta(1)) and TGF-beta(2) promote glioma cell migration via up-regulation of alpha(V) beta(3) integrin expression. Biochem Biophys Res Commun. 2000;268(2):607–11. http://dx.doi.org/10.1006/bbrc.2000.2176
  154. Liu S, Sun J, Lan Q. TGF-beta-induced miR10a/b expression promotes human glioma cell migration by targeting PTEN. Mol Med Rep. 2013;8(6):1741–6.
  155. Canazza A, Calatozzolo C, Fumagalli L, Bergantin A, Ghielmetti F, Fariselli L, et al. Increased migration of a human glioma cell line after in vitro CyberKnife irradiation. Cancer Biol Ther. 2011;12(7):629–33. http://dx.doi.org/10.4161/cbt.12.7.16862
  156. Thiery JP, Acloque H, Huang RY, Nieto MA. Epithelial-mesenchymal transitions in development and disease. Cell. 2009;139(5):871–90. http://dx.doi.org/10.1016/j.cell.2009.11.007
  157. Chaffer CL, Weinberg RA. A perspective on cancer cell metastasis. Science. 2011;331(6024): 1559–64. http://dx.doi.org/10.1126/science.1203543
  158. Iwadate Y. Epithelial-mesenchymal transition in glioblastoma progression. Oncol Lett. 2016;11(3):1615–20. http://dx.doi.org/10.3892/ol.2016.4113
  159. Moustakas A, Heldin CH. Signaling networks guiding epithelial-mesenchymal transitions during embryogenesis and cancer progression. Cancer Sci. 2007;98(10):1512–20. http://dx.doi.org/10.1111/j.1349-7006.2007.00550.x
  160. Khan MA, Chen HC, Zhang D, Fu J. Twist: A molecular target in cancer therapeutics. Tumour Biol. 2013;34(5):2497–506. http://dx.doi.org/10.1007/s13277-013-1002-x
  161. Elias MC, Tozer KR, Silber JR, Mikheeva S, Deng M, Morrison RS, et al. TWIST is expressed in human gliomas and promotes invasion. Neoplasia. 2005;7(9):824–37. http://dx.doi.org/10.1593/neo.04352
  162. Nordfors K, Haapasalo J, Makela K, Granberg KJ, Nykter M, Korja M, et al. Twist predicts poor outcome of patients with astrocytic glioma. J Clin Pathol. 2015;68(11):905–12. http://dx.doi.org/10.1136/jclinpath-2015-202868
  163. Wang Y, Shi J, Chai K, Ying X, Zhou BP. The role of snail in EMT and tumorigenesis. Curr Cancer Drug Targets. 2013;13(9):963–72. http://dx.doi.org/10.2174/15680096113136660102
  164. Han SP, Kim JH, Han ME, Sim HE, Kim KS, Yoon S, et al. SNAI1 is involved in the proliferation and migration of glioblastoma cells. Cell Mol Neurobiol. 2011;31(3):489–96. http://dx.doi.org/10.1007/s10571-010-9643-4
  165. Myung JK, Choi SA, Kim SK, Wang KC, Park SH. Snail plays an oncogenic role in glioblastoma by promoting epithelial mesenchymal transition. Int J Clin Exp Pathol. 2014;7(5):1977–87.
  166. Liu B, Dong H, Lin X, Yang X, Yue X, Yang J, et al. RND3 promotes Snail 1 protein degradation and inhibits glioblastoma cell migration and invasion. Oncotarget. 2016;7(50):82411–23. http://dx.doi.org/10.18632/oncotarget.12396
  167. Yang HW, Menon LG, Black PM, Carroll RS, Johnson MD. SNAI2/Slug promotes growth and invasion in human gliomas. BMC Cancer. 2010;10:301. http://dx.doi.org/10.1186/1471-2407-10-301
  168. Hill L, Browne G, Tulchinsky E. ZEB/miR-200 feedback loop: At the crossroads of signal transduction in cancer. Int J Cancer. 2013;132(4):745–54. http://dx.doi.org/10.1002/ijc.27708
  169. Wellner U, Schubert J, Burk UC, Schmalhofer O, Zhu F, Sonntag A, et al. The EMT-activator ZEB1 promotes tumorigenicity by repressing stemness-inhibiting microRNAs. Nat Cell Biol. 2009;11(12):1487–95. http://dx.doi.org/10.1038/ncb1998
  170. Zhang L, Zhang W, Li Y, Alvarez A, Li Z, Wang Y, et al. SHP-2-upregulated ZEB1 is important for PDGFRalpha-driven glioma epithelial-mesenchymal transition and invasion in mice and humans. Oncogene. 2016;35(43):5641–52. http://dx.doi.org/10.1038/onc.2016.100
  171. Qi S, Song Y, Peng Y, Wang H, Long H, Yu X, et al. ZEB2 mediates multiple pathways regulating cell proliferation, migration, invasion, and apoptosis in glioma. PLoS One. 2012;7(6):e38842. http://dx.doi.org/10.1371/journal.pone.0038842
  172. Maitre JL, Heisenberg CP. Three functions of cadherins in cell adhesion. Curr Biol. 2013;23(14): R626–33. http://dx.doi.org/10.1016/j.cub.2013.06.019
  173. Thiery JP. Epithelial-mesenchymal transitions in tumour progression. Nat Rev Cancer. 2002;2(6): 442–54. http://dx.doi.org/10.1038/nrc822
  174. Asano K, Kubo O, Tajika Y, Takakura K, Suzuki S. Expression of cadherin and CSF dissemination in malignant astrocytic tumors. Neurosurg Rev. 2000;23(1):39–44.
  175. Redies C. Cadherins in the central nervous system. Prog Neurobiol. 2000;61(6):611–48. http://dx.doi.org/10.1016/S0301-0082(99)00070-2
  176. Lewis-Tuffin LJ, Rodriguez F, Giannini C, Scheithauer B, Necela BM, Sarkaria JN, et al. Misregulated E-cadherin expression associated with an aggressive brain tumor phenotype. PLoS One. 2010;5(10):e13665. http://dx.doi.org/10.1371/journal.pone.0013665
  177. Camand E, Peglion F, Osmani N, Sanson M, Etienne-Manneville S. N-cadherin expression level modulates integrin-mediated polarity and strongly impacts on the speed and directionality of glial cell migration. J Cell Sci. 2012;125(Pt 4):844–57. http://dx.doi.org/10.1242/jcs.087668
  178. Musumeci G, Magro G, Cardile V, Coco M, Marzagalli R, Castrogiovanni P, et al. Characterization of matrix metalloproteinase-2 and -9, ADAM-10 and N-cadherin expression in human glioblastoma multiforme. Cell Tissue Res. 2015;362(1):45–60. http://dx.doi.org/10.1007/s00441-015-2197-5
  179. Asano K, Duntsch CD, Zhou Q, Weimar JD, Bordelon D, Robertson JH, et al. Correlation of N-cadherin expression in high grade gliomas with tissue invasion. J Neurooncol. 2004;70(1):3–15. http://dx.doi.org/10.1023/B:NEON.0000040811.14908.f2
  180. Perego C, Vanoni C, Massari S, Raimondi A, Pola S, Cattaneo MG, et al. Invasive behaviour of glioblastoma cell lines is associated with altered organisation of the cadherin-catenin adhesion system. J Cell Sci. 2002;115(Pt 16):3331–40.
  181. Siebzehnrubl FA, Silver DJ, Tugertimur B, Deleyrolle LP, Siebzehnrubl D, Sarkisian MR, et al. The ZEB1 pathway links glioblastoma initiation, invasion and chemoresistance. EMBO Mol Med. 2013;5(8):1196–212. http://dx.doi.org/10.1002/emmm.201302827